You are on page 1of 274

FLUORESCENT LAMPS

Philips Technical library

FLUORESCENT LAMPS

Edited by W. Elenbaas

MACMillAN
English edition N.V. Philips' Gloeilampenfabrieken, Eindhoven, 1971
Softcover reprint of the hardcover 2nd edition 1971
First edition 1959
Reprinted 1962
Second edition 1971
All rights reserved. No part ofthis publication may be reproduced or transmitted,
in any form or by any means, without permission.

SBN 333 054172

ISBN 978-1-349-00363-1 ISBN 978-1-349-00361-7 (eBook)


DOI 10.1007/978-1-349-00361-7

First published in English by


THE MACMILLAN PRESS LTD
London and Basingstoke
Associated companies in New York, Toronto, Melbourne
Dublin, Johannesburg and Madras

PHILIPS
Trademarks of N.V. Philips' Gloeilampenfabrieken
Preface
The 1962 edition of 'Fluorescent lamps and lighting' has here been brought
up to date. Since the extent of the subject has again increased, it was decided
not to deal with applications in this book.
The authors of the first nine chapters of the 1962 book have been joined
by Messrs. Moerkens and Vrenken. The sequence of the chapters has been
ch~nged somewhat.
We start with a chapter on 'gaseous discharges' and one on 'the discharge
of the fluorescent lamp'. Then follows 'luminescence' and 'luminescent sub-
stances' after which a chapter 'lamp design and lamp manufacture' is in-
serted. A chapter on 'colour and colour rendering' treats this subject, which
is so closely connected with the phosphors that we like to treat it here. The
'stabilisation' and the 'lamp types and circuits' are treated in the two next
chapters, after which chapters on 'invertors and convertors', on 'dimmers'
and on 'balast design' follow. The book is concluded with a chapter on
'installations' in which regulations, radio interference, heat problems, etc.
are treated. Electronics have become very important in the circuits used and
have therefore been given ample attention in this book.
The sections marked at the beginning and end with a dagger (t) are in-
tended for those readers who are interested in a more detailed discussion.
If desired, however, these sections can be passed over without disturbing
the continuity of the remainder of the text.
W. ELENBAAS
Table of Contents
Preface
1 Gaseous discharges W. Elenbaas 1
Introduction - Electron emissiOn - Structure of the atom and
mechanism of radiation and ionisation in gaseous discharges - Elastic
and inelastic collisions - The potential gradient of the discharge -
The current-voltage characteristic - Stabilisation - Ignition

2 The discharge of the fluorescent lamp W. Elenbaas 19


Why the low pressure mercury vapour discharge? - Lamp dimen-
sions- How much light can be expected from a fluorescent lamp? -
The energy balance of the fluorescent lamp - The efficiency as a
function of different parameters

3 Luminescence, fluorescence and phosphorescence


J. L. Ouweltjes 32
Incandescence and luminescence - Some practical aspects of the
luminescence of solids - The emission and absorption spectra of
solid substances - Some further considerations of the electron tran-
sitions involved in luminescence

4 Luminescent substances J. L. Ouweltjes 41


Chemical composition of phosphors- Requirements for the practical
application of phosphors - Physical properties of phosphors - Pre-
paration of phosphors

5 Lamp design and lamp manufacture L. E. Vrenken 52


Essential parts- Bulbs- Electrodes- Mercury and rare gas- Caps
- Lamp making - Coating the tube wall with a phosphor -Processing
in the lehr - Processing on the exhaust machine - Activation of the
electrodes

6 Colour and colour rendering


A. A. Kruithof, J. L. Ouweltjes 71
Blending of fluorescent materials - The fundamentals of colour
vision - Construction of the chromaticity diagram - Computing
colour points - Application of chromatics in the development of
fluorescent lamps - Choice of colours for general lighting fluorescent
lamps - Colour tolerances - Colour rendering- The specification of
the colour rendering properties - The shortcomings of the 'standard'
lamps - Data for some of the Philips de Luxe lamps
7 Stabilisation of the discharge Th. Hehenkamp104
Introduction - Direct-current supply - A.C. operation - Conse-
quences of current distortion - Measurement of ballasts

8 Lamp types and circuits J. Funke, J. C. Moerkens120


Introduction - Starter switches - Lamp types for switch start opera-
tion - Starter circuits - Lamp types for starterless circuits - Circuits
for preheat starterless lamps - Instant start circuits (cold starting) -
Lamp operation at higher frequencies -D.C. circuits and lamps -
Lamp types for special purposes

9 Inverters and converters Th. Hehenkamp181


Introduction - Principle of the transistor inverter - Inverter with
reduced switching losses - Cooling of the transistor - Applications
of transistor inverters - The thyristor as a switch - Inverter with
forced commutation - Self-commutating inverters - Applications of
thyristor inverters and converters

10 Dimming of fluorescent lamps J. C. Moerkens198


Introduction - The principle of dimming fluorescent lamps -
Influence of fluctuations in mains voltage - Control circuit with
stabilising effect - Symmetry of the lamp current - Controlling high
powers (central control unit) - Outdoor lighting (influence of the
temperature) - Automatic control - Auxiliary equipment - Dimmer
for low powers - Dimmer ballasts

11 Ballast Design Th. Hehenkamplll


Introduction - Iron circuits - Different types of core constructions
- Choke dimensions in relation to losses and insulation temperature
- Copper space factor - Relation between insulation temperature and
life - Cooling of the ballast - Ballasts of small cross-section - Con-
struction of capacitors- Capacitor life- Ballast noise- Noise mea-
surement

121nstallations J. Funke, J. C. Moerkens245


Temperature problems with ballasts - Radio interference- Higher
harmonics in the mains current - Lamp performance at low tem-
perature - Dealing with audio frequency signals in the mains
Chapter 1

Gaseous discharges
W. Elenbaas

1.1 Introduction
In the fluorescent lamp the transfer of electric energy into visible light
takes place in two steps. First the electric energy is partly transferred into
invisible ultra-violet radiation mainly of wavelength 2537 nm. The amount
of visible radiation produced by the discharge itself is small compared with
the ultra-violet radiation.
The ultra-violet radiation produced by the discharge falls on the fluores-
cent powder, which is situated at the inner wall of the discharge tube and
is there transferred into visible light (in special lamps mainly invisible ra-
diation is produced for special purposes). The production of radiation by
the discharge is thus the first step in the light production and we will there-
fore start with a general treatment of gas discharges.
1.2 Electron emission
In gaseous discharges, charged particles - namely electrons and positive
ions - move in a gas between two electrodes. Negative ions occur in some
types of discharge, but these need not be considered here. Discharges without
electrodes in the gas (possible at high frequencies) are also not treated in
this book.
2 FLUORESCENT LAMPS

+
Fig. 1.1. Gaseous discharge tube in
series with a D.C. voltage V2 and a
resistor R. The voltage V 1 serves to
heat the cathode.

Let us first look at the arrangement illustrated in Fig. 1.1, which shows a
lamp connected to a source of direct current; when the lamp operates, a
potential difference V, ( = V2 - IR) exists between the electrodes. On
average, the electrons travel from the negative electrode (cathode) to the
positive electrode (anode), whilst the positive ions, whose average rate of
progress is much slower than that of the electrons, move from anode to
cathode.
Since the average movement of the electrons is towards the anode, there
must be a continuous supply of electrons by or near the cathode. The pro-
cess of electron supply by the cathode is known as electron emission. More-
over, positive ions and electrons are created throughout the whole discharge
(see Sections 1.3 and 1.4).
The emission of electrons is a very important feature of a gaseous dis-
charge; so much so, that the manner of the emission determines whether
the discharge will be called a glow discharge or an arc discharge. By glow
discharge is meant the discharge from a cold cathode, and by an arc dis-
charge that from a hot cathode *.
Free electrons occurring in the metal of the cathode are not normally able
to emerge from the cathode into the surrounding medium; to make this
possible the electron requires a certain minimum amount of energy; this
energy is expressed in terms of the electron-volt, defined as the energy
acquired by an electron in changing its potential by 1 volt. If, for a given
metal, this minimum energy be q; electron-volts, q; is called the thermionic
work function of the metal.
The mechanical analogy of these electrical forces, which prevent the
electrons from leaving the metal is shown in Fig. 1.2. The box is filled with
marbles up to a level lying a distance h below the rim p of the box. Unless
some force acts upon them, the marbles cannot rise to the level p; to reach
this level a marble in the uppermost row must acquire an energy of at least
mgh (m = mass of the marbles and g = acceleration due to gravity). The
box may be taken to represent the metal in which the electrons are located,
whilst the rim is analogous to the space outside the metal to which the

p p Fig. 1.2. In the same way that energy is


required to move the marbles out of the
box, energy must be imparted to the
electrons to enable them to leave the
metal.

* Arc discharges with cold electrodes, as occur on liquid mercury or on copper electrodes
are left out of consideration. Here the cathode fall is also small (approximately 10 V)
which is characteristic for the arc discharge, whereas the cathode fall of the glow dis-
charge equals some 100 V.
GASEOUS DISCHARGES 3
electrons cannot pass without energy being imparted to them. The dis-
tance h and the quantity cp are wholly analogous; the difference is however
that the electrons move within the metal, whereas the marbles are stationary
in the box.
Let us now consider two ways in which the electrons can emerge from
the metal.

1.2.1 Arc discharge


When a metal is heated, the mean velocity of the electrons in the metal is
increased. The velocity for all electrons at any given moment is, however,
not the same. There exists a certain distribution of velocities covering both
slow and rapidly moving electrons, and those of the latter which move
quickly enough and moreover strike the surface are able to leave the metal.
If, furthermore, an electric field exists in the adjoining space and this field
is so oriented as to exert on the electrons a force in a direction away from
the metal, these electrons may move still further away from the metal. lf
the temperature of the metal is increased, more and more electrons acquire
a sufficient velocity to escape from the metal, and the emission thus increases.
Fig. 1.1 therefore includes a voltage V 1 to maintain the cathode at an
elevated temperature. In Fig. 1.2 it is obviously an advantage if the dis-
tance h is small, so that marbles whose kinetic energies are lower can also
reach the rim; in the case of electron emission it is thus advisable to employ
a substance having a low work function, and, in practice, this is achieved
by covering the cathode with a thin layer of an alkaline earth oxide such as
BaO, SrO or CaO or a mixture of these.

1.2.2 Glow discharge


Suppose that a projectile is shot into the box of marbles shown in Fig. 1.2.
One or more of the latter may then be ejected from the box, depending
naturally on the nature of the projectile, its direction and velocity, and the
distance h. In a glow discharge the projectile takes the form of a positive
ion, which is accelerated just in front of the cathode by reason of the electric
field in that region, and then alights on the cathode. By no means every
positive ion arriving at the cathode liberates an electron; the probability of
this happening depends on the ion velocity and also on the work function.
In actual fact it appears that the odds are about 1 to 100, that is to say,
for every 100 positive ions arriving, only about one liberates an extra elec-
tron (one electron being used to neutralise the positive ion). For this kind
of emission the cathode need not be heated at all; however, its temperature
will increase somewhat owing to the bombardment of the positive ions, but
this temperature rise is not essential to the emission process. In this case,
therefore, we speak of cold-cathode emission. The electron emission per
square centimeter is very much lower than in the case of the arc discharge.
For the same current, thus, the cathode of the glow discharge must be con-
siderably larger than that of an arc, and, in fact, it usually consists of a
sheet-iron cylinder. In order to maintain the advantage of a low work
function and thus enhance the chances of liberating electrons, these elec-
trodes are sometimes coated with alkaline earth oxides. In spite of this, the
4 FLUORESCENT LAMPS

currents used in glow discharges are mostly much smaller than those in
arcs.
One essential feature of the glow discharge is a high potential drop just
in front of the cathode, to ensure that the velocity of the positive ions arriving
at the cathode will be sufficiently high. The potential drop (the cathode fall)
adjusts itself automatically so that the electron emission acquires the right
value. The cathode fall depends on the nature of the cathode surface (this
determines the value of cp) and on the gas (this determines the kind of
projectile). The cathode fall of the glow discharge is some 100 V, as against
only about 10 V for the arc discharge.
1.3 Structure of the atom and mechanism of radiation and ioni-
sation in gaseous discharges
According to Bohr, based on experiments of Rutherford, the atom may
be considered as consisting of a very small nucleus in which all the mass is
concentrated and which has a positive charge, with a number of electrons
circling around it in the same way as the planets circle around the sun. The
attractive force in the case of the atom is of an electric nature due to the
positive charge of the nucleus and the negative charge of the electron. In a
normal atom the number of electrons associated with it is just sufficient for
their combined negative charge to counterbalance the positive charge of the
nucleus as far as any external effect is concerned. The difference between
the elements is that the nuclear charge and therefore also the number of
revolving electrons increases from 1 for hydrogen to 92 for uranium (and
still higher for artificial elements). The nuclear charge of mercury, for
instance, is +80 'electron charges', and 80 electrons revolve around the
nucleus. Now, these electrons describe certain definite orbits only; inter-
mediate paths do not occur. The electron occupying the orbit farthest
removed from the nucleus is the least influenced by the nucleus, because of
the larger distance, and also since the other electrons pass between it and
the nucleus and thus screen it to a considerable extent from the electric field
of the nucleus. This remote electron is therefore less strongly associated
with the system and, in gaseous discharges, the energy of any colliding elec-
tron is generally so small that only this remote electron can be affected and
be thrown into an orbit other than its normal one. In order to bring an
electron into one of these more remote orbits, energy has to be applied to
it, in the same way that energy is required to lift objects against the force
of gravity. In the case of the atom, however, only certain specific higher
orbits are involved, without the possibility of any intermediate orbits; this
can best be illustrated as in Fig. 1.3, which depicts various energy levels of
the atom. The zero line represents the energy of the atom with all the elec-
trons in their normal orbits (ground state). The unit in the energy scale is
the electron-volt, i.e. the energy change of an electron involved in changing
its potential by 1 V (1 electron-volt= 1 eV = 16 x IQ- 19 joule). If the outer
electron be removed to a higher orbit, the energy of the atom may be repre-
sented by level A in the diagram; the horizontal distance from the energy
axis is here of no significance, in contrast to Fig. 1.4, which is an energy-
level diagram for mercury, where the levels are also grouped horizontally,
to make allowance for certain characteristics of these levels which make it
GASEOUS DISCHARGES 5

eV;~~
E ----8
t
Fig. 1.3. Diagram showing the energy
levels of an atom. The zero line repre- ev,. [ -A
sents the atom in the normal state. The
atom is in the 'excited' state at A or B, 0
whilst the hatched zone denotes the
state of ionisation.
~-------------

desirable to plot them in series; in Fig. 1.4 levels belonging to the same
series appear one above the other. They are denoted by letters and numbers
(alongside each level) which are indicative of the characteristics mentioned
above.
Returning to Fig. 1.3, the atom may be in the normal condition (the ground
state), or in condition A orB, or in a number of other conditions not repre-
sented, but which depend on the particular orbit in which the outer electron
is located. This electron can also become completely detached from the rest
of the atom, this process involving a minimum amount of energy equal to
eV; (V; is the ionisation potential). The condition represented by level C
is such that the electron is dissociated from the atom and moves at a velo-
city v whereby its kinetic energy t mv 2 is equal to eV1 As the velocity of
the electron may vary continuously, the conditions at C cover the whole
of the hatched zone (and above) in Fig. 1.3, in contrast with the discrete
levels below V;.
Collisions between atoms in the normal state and an electron whose
kinetic energy t mv 2 (v =velocity with respect to the atom) is less than eV,
(A representing the lowest excited state) can be only of the 'elastic' kind
(see Section 1.4). If the energy is greater thane V, however, collision involves
the probability that the atom, whilst absorbing the amount of energy e V,
will be elevated to condition A, whereas the colliding electron goes on its
way with an energy which is decreased by the amount e V,. The same holds
for the other levels. However, the atom does not remain long in the newly
acquired excited state, returning spontaneously after a very short space of
time (of the order of w-s s) to its normal state. In doing so, it emits the
energy e V, in the form of radiation. The frequency of this radiation is directly
proportional to the energy change ilE. Thus:
ilE = e l1 V = hv =he/A. (1.1)
where l1 Vis the distance between the two levels, v the frequency, h Planck's
constant, c the velocity of light and A. the wavelength. Substitution of the
values for e, h and c in equation 1.1 gives us:
A.il v = 1 239 (1.2)
with V in volts and A. in nm. As we are concerned with discrete energy
levels A, B and so on, so that l1 V can assume certain values only, the radia-
6 FLUORESCENT LAMPS

tion occurs only at certain wavelengths; that is to say, the spectrum is not
continuous but consists of a number of spectral lines corresponding to the
various differences between the energy levels. The wavelengths of the mer-
cury lines can be ascertained from Fig. 1.4; level 23 P 1 occurs at 488 V,
so that the wavelength corresponding to the transition from 2 3 P 1 to the
ground level is 1 239/488 = 254 nm.

'~.....---9.55
3 381 ~9.22
8.84 \
334

fl So--"---------------------'
Fig. 1.4. Energy level diagram of the mercury atom showing the more important lines.
The distance of the levels to the ground level 1 1 S 0 is given in electron-volts and the
wavelengths corresponding to the transitions from one level to another are indicated
in nm.

Before returning to its original level and emitting light in the process, it
is also possible for an atom being in the condition A (Fig. 1.3) to collide
with a second electron and pass from condition A to condition B, or to the
ionised state (cumulative excitation or cumulative ionisation), or the exci-
tation energy may be re-converted into kinetic energy of the colliding elec-
tron, the atom then returning to its normal state. The cumulative excitation
and ionisation will occur more frequently when the lifetime of the condi-
tion A is longer. The duration of the excited condition, stated earlier to be
GASEOUS DISCHARGES 7
about 1o- s, does not actually apply to all levels: there are certain levels,
8

the so-called meta-stable levels, in which the atom, if further undisturbed,


remains for a longer period (up to 01 s); in other words the transitions
between the different levels are not equally probable, the transition probability
if the initial level is metastable, being very small. In Fig. 1.4 for example,
the 23P 2 and 2 3P 0 levels are meta-stable; these meta-stable levels sometimes
play a very important part in gaseous discharges, because they give rise to
cumulative excitation and ionisation, or because this excitation energy is
used to ionise another kind of atom (Penning effect, see Section 1.8.2).

1.4 Elastic and inelastic collisions


The path of an electron will now be traced through the tube after leaving
the cathode in the manner described in Section 1.2.1 or 1.2.2.
Between the electrodes (Fig. 1.1) there is an electric field which accelerates
the electron in the direction cathode-anode. If no gas is present in the tube,
the electron, under the influence of the field, describes a parabola (or a
straight line if the electron emerges from the cathode exactly in the direction
of the field), and will be hurled against the anode and probably disappear
within it. In this case the energy that it has acquired on its way from the
cathode to the anode, plus an amount of energy corresponding to the work
function of the anode material (the entry of the electron into the metal
liberates q; electron-volts) will be converted into heat in the anode. This,
however, is not the object for which the lamp is designed. We therefore intro-
duce a gas between the electrodes, so that the electrons can no longer follow
their parabolic paths without hindrance. The higher the concentration of the
gas, the greater are the chances that the electron will collide with a gas atom
and thus be thrown off its original course. Since the electron is considerably
lighter than the gas atom (for instance approx. 370 000 times lighter than
the Hg atom), the latter is practically unaffected by the collision. The elec-
tron, on the other hand, changes direction with every collision and thus
follows a zig-zag path which is in the general direction of the anode, but
which is much longer than the direct path from cathode to anode (up to
several hundred times longer). The number of collisions that the electron
experiences is very high; in fact, at a velocity corresponding to 1 eV in a
gas at a pressure of 1 torr, the electron collides some 10 9 times per second.
Even though the transfer of energy accompanying an elastic collision between
an electron and an atom is only small, the total amount of energy transferred
per second in such a large number of collisions is not negligible. In conse-
quence the average velocity of the gas atoms gradually increases, which
means that the temperature of the gas rises. The gas temperature will rise
until the energy lost by thermal heat dissipation equals the energy acquired
by the collisions. In this way electrical energy is converted into heat and is
lost for the purpose of generation of light. As these losses occur in the gas
they are referred to as gas losses or volume losses; since the number of
electron collisions is proportional to the number of atoms per cubic centi-
metre, these losses are approximately proportional to the gas pressure. Since
8 FLUORESCENT LAMPS
the energy transfer accompanying a collision is larger if the mass of the atom
is smaller, the gas losses increase with decreasing atomic weight of the gas.
In addition to these so-called elastic collisions between electrons and atoms,
which leave the structure of the atom unaffected, it is possible for the atom
to undergo a radical change when the velocity of the electron is high enough.
As stated in Section 1.3 the atom consists of a nucleus around which
numerous electrons rotate. Now, the colliding electron may knock an atom-
electron (of the outer orbit) into an orbit which is more remote from the
nucleus, thus producing what is known as an excited atom. Generally
speaking, the atom will quickly (after about w-s s) return by itself to its
original state, and by doing so radiates energy, which is the desired process.
A further possibility is that, as a result of the collision, the atom-electron
may become completely dissociated from the atom itself, in which case we
say that the atom is ionised; it is then one electron short, thus has a positive
charge and is therefore called a positive ion. The liberated electron will be
accelerated by the electric field and may in its turn, after several elastic
collisions, excite or ionise another atom. This ionisation process is essential
to the continuance of the discharge, since otherwise only electrons emitted
by the cathode would pass down the tube and there would be a complete
absence of positive ions, resulting in heavy negative space charges and
intense electric fields. Moreover, electrons arriving at the wall of the tube
are lost and have to be replaced by the process of ionisation in the gas.
The ionising collisions are therefore indispensable for the discharge.

1.5 The potential gradient of the discharge


The potential gradient between cathode and anode is not uniform all the
way down the tube. The electric field is constant only in the attenuated part
of the discharge, that is, in the positive column where the electrodes have no
further effect on it and where the potential is a linear function of the distance
from the electrodes. In that zone there is, practically speaking, no space
charge; wherever a space charge occurs (and this is usually the case in the
vicinity of the electrodes), the electric field is not constant. There is often a
positive space charge just in front of the cathode and the field strength in
that region is consequently high. In a glow discharge the resultant potential
difference over the cathode fall area is of the order of 100 V.
The increased electric field arising from a space charge is demonstrated
in Fig. 1.5. In the absence of a space charge the electric lines of force run
directly from cathode to anode (Fig. 1.5a); the field E is then the same at
all points (Fig. 1.5b) and the potential V increases linearly from cathode to
anode (Fig. 1.5c). When a positive space charge occurs, however, lines of
force emanating from the cathode terminate at the positive space charge
(Fig. 1.5d). The flux density and therefore also the electric field are then
greater in front of the cathode (Fig. 1.5e). Fig. 1.5f shows the course of the
potential. Sometimes a negative charge will occur in the region of the anode,
so that the electric field at that point also exceeds that of the positive column,
and this results in an anode fall as well. This is reduced, however, when the
GASEOUS DISCHARGES 9

Fig. 1.5. Lines of force (a, d and g), electric field E (b, e and h) and potential V (c, f
and i) without space charge (a, b and c), with positive space charge in front of the
cathode (d, e andf) and with positive space charge in front of the cathode plus negative
space charge in front of the anode (g, h and i). I = distance to cathode; v. = anode
fall; Vc = cathode fall; c = positive column.

current density is decreased, and the anode fall may even drop to zero if the
anode itself is sufficiently large. Figs 1.5g, hand i illustrate the lines of force,
the electric field and the potential in the case of a positive space charge in
front of the cathode and a negative space charge in front of the anode.

1.6 The current-voltage characteristic


The characteristic of a discharge represents the relationship between the
current passing through the lamp and the voltage across it, after the manner
shown by curve a in Fig. 1.6. Generally speaking, the voltage across an arc

l1
f111 --JviA
r ~--
1 I
I I
I I
I I
I I
1 I
Fig. 1.6. Diagram showing the voltage I..:~!, I
across and arc discharge as a function
of the current. 0 It
10 FLUORESCENT LAMPS

discharge drops when the current rises, resulting in what is termed a negative
characteristic, although this does not necessarily apply to all arc discharges.
The characteristic of the fluorescent lamp is in effect a combination of the
characteristic of the cathode fall, the characteristic of the elongated part of
the discharge (the positive column), which in a fluorescent lamp stretches
from within a few centimetres of the cathode to a point near the anode,
plus the characteristic of the anode fall.

Vc
t A

Fig. 1.7. Potential difference Vc over the cathode fall area in a wide range of current values
(the points A in Figs 1.6 and 1. 7 represent the same current).
1= normal glow discharge
2 = abnormal glow discharge
3 = arc discharge

Fig. 1. 7 shows the characteristic of the cathode fall of a not externally


heated cathode, plotted on a logarithmic current-scale, in order to bring
out the details in the low current part of the range. In a normal glow dis-
charge the cathode fall is independent of the current; the cathode in this
low-current region is not wholly covered by light, and the covered area is
proportional to the current, so that the current density at the cathode
remains constant. With increasing current, these conditions only remain
possible until the whole cathode area becomes luminous. If the current
increases any further, the current density at the cathode has to increase and
this increase is accompanied by a rise in voltage which brings us into the
range of the abnormal glow discharge, in which zone the glow switches work
(reviewed in Section 8.2). Here, then, the voltage increases with the current,
and the characteristic is positive. The increase in current as well as in cathode
fall causes the amount of heat generated at the cathode to increase (more
positive ions striking the cathode at higher and higher velocities), and the
temperature of the cathode itself rises. At the point P in Fig. 1. 7 the tem-
perature is in fact so high that the cathode commences to emit thermionically.
GASEOUS DISCHARGES 11
A further increase in current results in a still higher cathode temperature, as
a result of which the emission of the cathode increases and the voltage drops.
From here onwards we are dealing with an arc discharge, to which type the
hot-cathode fluorescent lamp belongs. For this lamp the cathode fall is of
the order of 10 V.
The anode fall is small and does not vary a great deal, so that it is not very
interesting for this discussion. On the other hand, the characteristic of the
positive column is important, especially in long tubes, since it accounts for
a considerable portion of the total voltage. The characteristic of the positive
column of the fluorescent lamp is negative. This is due to the fact that, with
increasing current, cumulative ionisation occurs more and more, since the
concentration of excited atoms becomes greater and the chances of electrons
colliding with an excited atom are thus enhanced. Suppose for a moment
that cumulative ionisation and excitation did not occur; in that event we
would expect the voltage of the positive column to be independent of the
current. This can be explained as follows: at twice the current, and the same
gradient, twice the number of electrons are moving. With the same gradient
in the positive column, the velocity distribution is the same, so that twice
as many ionisations will occur, thus doubling the production of ions and
electrons. If the loss of charged particles is proportional to the concentration,
then the concentration under equilibrium conditions will be proportional to
the production. Thus, at double the current, the concentration of electrons
and positive ions is also twice as large and the double current can flow at
the same gradient, which is the supposition from which we started. If, how-
ever, 100% cumulative ionisation occurs, the frequency of ionisation at twice
the current (and the same electron velocity distribution) will be four times
as high; the electron concentration is then four-fold and the potential gra-
dient required is only half. Since with the smaller potential gradient the
velocity of the electrons will diminish, the ionisation frequency will be less
than 4 times and the potential gradient will be somewhat more than half.
If only part of the ionisation is cumulative, the voltage will drop to an even
lesser extent.
Thus the characteristic of the positive column of a fluorescent lamp is
negative and the characteristic of the lamp voltage as a whole (the sum of
the cathode fall, the anode fall and the positive column voltage) is also
negative, in accordance with Fig. 1.6.

1.7 Stabilisation
As explained in the previous section, the characteristic of the fluorescent
lamp is negative. This causes the lamp to fail if connected directly to a
voltage supply high enough to produce ignition. To understand this let us
suppose that the lamp is made to ignite on an open voltage or a mains
voltage V0 (see Fig. 1.6) and that the current is / 1 (point A); this means that
the full voltage is applied across the lamp and that open voltage V0 and lamp
voltage V1 are in equilibrium. This equilibrium, however, is unstable. If the
current momentarily increases by an amount AI, the lamp voltage in the
12 FLUORESCENT LAMPS

equilibrium condition would drop to the extent of Ll V. However, the full


open voltage V0 remains applied across the lamp. The electric field strength
in the discharge path is consequently greater than is necessary for a current
I 1 + LII and the electrons are accelerated to a greater extent than neces-
sary for this current, so that the latter increases and continues to build up
until either the tube bursts or the mains fuse blows. For this reason dis-
charge lamps are operated in series with a choke (on a.c.), or with a resistor.
Connection with a capacitor in series does not give good results, at least at
50 Hz (see Section 2.5.6 and Section 7.3.3), but combinations of chokes and
capacitors are often employed, as explained in Chapter 8.

D__ ----

Fig. 1.8. Diagram showing sta-


bilisation of a discharge (charac-
teristic a) by means of a re-
I sistor (characteristic b).

Let us first take the simplest case of a lamp operating on direct current,
stabilisation being by means of a resistor. The characteristic of the discharge
is represented by the curve a in Fig. 1.8. Let the open voltage be V0 ; when
the current I flows, the voltage drop across the resistor is IR, and in Fig. 1.8
this voltage is depicted as a function of I as a line drawn downwards from
the dotted horizontal line V0 D. Assuming that R is constant (independent
of I), the course of IR as a function of I will be represented by the straight
line b. It is essential for equilibrium that the sum of the voltage across the
resistor (IR) and that across the lamp (V1) exactly equals the open voltage V0 ,
and this condition is satisfied at points A and B. We have now to investigate
whether these equilibria are stable or unstable. With regard to A, let us
suppose once again that the current increases by the amount LI I. This brings
us to point A ' , leaving the voltage Ll V in hand, which means that the current,
as explained above, again increases, taking us still further from the point A .
This shows that A is a point of unstable equilibrium. In the case of B, if
GASEOUS DISCHARGES 13
the current increases to / 2 + LJI, there is a potential CD across the resis-
tor leaving only CG available for the discharge whereas for equilibrium,
B'G would be required. The field strength is thus smaller than the value
needed to maintain the current / 2 + LJI. The latter decreases therefore, and
returns to B; hence B is a point of stable equilibrium.
Proceeding along these lines it will at once be clear that, if the value of
the resistor be increased or the open voltage reduced, the lamp must, at a
given moment, extinguish because the line b no longer intersects the line a.
Suppose the open potential V0 to be constant and R variable. When R in-

Fig. 1.9. Effect of variation of


the series resistor and of the
mains voltage upon stabilisa-
tion.

creases, the line b rotates around the point P in the direction of the arrow
in Fig. 1.9; the point of equilibrium then moves from B in the direction of
the double arrow, but only as far as B'. If the current is reduced any further,
by increasing the resistance R, the point of intersection disappears and the
lamp is extinguished. It is possible, however, to reduce the current by using
a higher mains voltage, V0 ' (with higher R), as shown in Fig. 1.9, where
the conditions between B' and B" can be realised by means of a mains
voltage V0 '.
We finally consider stabilisation by means of a choke, this being the most
commonly used method for lamps operated on alternating current.
Here the situation is very analogous to that described above, except that
in this case the sum of the voltage across the lamp and that across the
ballast is not equal to the open voltage, but approximately Vo 2 = Vl 2 + VL 2 '
where VL is the voltage across the choke (for the exact equation see Section
7.3).
14 FLUORESCENT LAMPS

~2
0

Fig. 1.1 0. Stabilisation in the case of


alternating current by means of a
choke.

In Fig. 1.10 the square of the voltage is shown plotted as a function of I,


with VL 2 drawn vertically downwards from the dotted horizontal line at
height V0 2 (curve b). The point of intersection, B, of this line with the char-
acteristic a of the discharge once more represents the point at which the
current becomes stable.
The way in which stabilisation is carried out in practice, is described in
Chapter 7.

1.8 Ignition
The foregoing considerations are based on the pre-supposition that the
lamp will actually light. In many cases, however, if the lamp were to be
connected in series with a resistor or choke to a 220 V a.c. mains the lamp
would not light. In fact, a very much higher voltage has to be applied, or
other devices must be resorted to, before ignition can take place. After
ignition the potential across the lamp will drop from that of the applied open
voltage to the arc voltage of the lamp, at which level the current is deter-
mined by the characteristic of the lamp and the resistor as shown in Section
1.7.
1.8.1 Starting methods
The methods employed to start fluorescent lamps are described below.
One method is to operate the lamp on a transformer, the open voltage of
which is sufficiently high to ignite the lamp. In order to understand more
easily what is happening, we will assume that a d.c. voltage of increasing
magnitude is applied to the lamp. At low voltage and with the cathode at
room temperature, very few electrons are emitted and nothing further hap-
pens; the few electrons liberated by cosmic radiation in the tube are not
sufficiently accelerated to cause ionisation. If the applied voltage is increased,
GASEOUS DISCHARGES 15
however, a stage is reached at which the higher electric field can result in
ionisation by the electrons present in the tube. Positive ions then proceed
in the direction of the cathode and, striking the latter, may liberate an elec-
tron from it, as seen in Section 1.2.2. Now, for what further processes would
such an electron be responsible? It travels more or less at random but in the
general direction toward the anode and may produce ionisation; when this
happens, the resultant ion and also the old electron and the new electron fol-
low their respective paths. Both electrons may now ionise etc. The positive
ions, all of which are produced in consequence of the one electron under con-
sideration, travel towards the cathode. Some of them reach the latter and if
together they liberate one other electron from the cathode, the discharge
is maintained. Ignition is thus made easier when the voltage applied to the
lamp is increased (since ionisation takes place more readily at higher elec-
tric fields) and also when the electrode is able to emit electrons more
readily. We have already seen that this is the case when the work function
is low, but the type, as well as the pressure, of the gas will also affect the
result, since, apart from the fact that they determine the ionisation rate, the
kind of positive ion striking the cathode is dependent on the type of gas. If
the glow discharge is established between the electrodes as described above
and the open circuit voltage and ballast are adequately chosen, the discharge
will develop into an arc with the required current (see also Fig. 1.7). This
kind of ignition occurs with so-called instant start fluorescent lamps where
the electrode is not preheated.
A method of assisting starting is to make use of a glow current other than
one flowing between both main electrodes. This auxiliary glow current may
be very weak if it is caused by capacitive action of metal parts outside the lamp,
or it may be larger when starting electrodes or stripes inside the lamp
are present. In both cases, in addition to the formation of charged particles,
the heating up of a main electrode by this glow discharge may be of extra
assistance.
A further method to improve starting is based on the phenomenon shown
in Fig. 1.11, where the starting voltage of a 40 W lamp is given as a function
of the electrical resistance of the tube surface *. Starting is thus easier with
either a very low or a very high resistance. The former may be achieved by
applying a conductive stripe along the tube and the latter by coating the
tube with a water-repellent layer, for instance a silicone film, which prevents
the formation of an uninterrupted water-film.
Although in the cases described above, the electrodes are not heated by a
current through the electrode coil, the temperature of the cathode neverthe-
less increases during the starting process and, after starting, the electrode
emits thermionically. We are dealing, thus, with an arc discharge.
In most cases, the cathode is heated before or during the time that the
voltage is applied across the lamp and thus production of electrons by im-
pinging positive ions as outlined above is unnecessary, since the cathode emits
electrons because of its high temperature. An example of the ignition poten-
tial of a given lamp as a function of the heating current through the electro-

R.N. Thayer and D.D. Hinman, Trans. Ill. Eng. Soc., 40, 641 (1945).
16 FLUORESCENT LAMPS

r\
800

7 \
700
l1gn

7 \
600

1 500
v \
400

300
v ~
-
200 6
10 1010 To"
--R(.II.)

Fig. 1.11. Starting voltage of instant start lamps (40 W) as a function of the longitudinal
electrical resistance R of the lamp surface. (R. N. Thayer and D. D. Hinman, Trans.
Ill. Eng. Soc, 40, 641, 1945).

des (thus also as a function of the temperature of the electrodes - upper


scale) is given in Fig. 1.12. This curve depends on the longitudinal resistance
of the tube wall (Fig. 1.11 ). The curve of Fig. 1.12 is measured on a lamp
with an outer stripe connected to one of the electrodes. Without this the
curve lies somewhat higher and the measurements are less reliable. The gene-
ral form is however the same. As shown in Fig. 1.12 the starting poten-

-rroo
300 UJO 500 600 700 800 900
30'U
oft

'1\
0

200

150

0.2 0.25 0.3 0.35 0.4


-I(AJ
Fig. 1.12. Starting potential of a 40 W fluorescent lamp with a connected outer stripe as
a function of the heating current through the electrodes. The upper scale gives the elec-
trode temperature corresponding to the heating current.
GASEOUS DISCHARGES 17
tial at a sufficiently high temperature is considerably less than with cold
electrodes. This pre-heating of the cathode can be carried out in different
ways, the diameter of the electrode wire and the value of the heating current
being so adjusted that the required electrode temperature is obtained. In some
cases the voltage over the electrode during this heating is made so high (by
choosing the resistance of the coil sufficiently high) that arcing parallel to
the electrode occurs before the main arc strikes. The electrons of this
parallel arc help to strike the main arc.
The methodes used in practice to preheat the electrodes are treated in
Chapter 8.

1.8.2 Influence of the lamp geometry and of the pressure and the nature of
the gas on starting
The starting potential is, of course, dependent on the dimensions of the
tube, and on the nature and pressure of the gas. Generally speaking ignition
is more difficult in long narrow tubes than in shorter, wider ones.
The effect of the gas pressure will be easily understood as follows: in a
vacuum, ionisation cannot take place at all and the breakdown voltage is
therefore infinitely high, whilst at very high pressures the average unimpeded
path of the electron is very short and the latter loses energy in undergoing
elastic collisions. The electron is therefore unable to acquire a sufficient veloc-
ity to cause ionisation and the ignition is, consequently, again very difficult.
It is therefore not surprising that the curve representing the ignition voltage
as a function of the gas pressure assumes the form shown in Fig. 1.13
(Paschen curve).

Fig. 1.13. Ignition voltage as


a function of the gas pressure
(diagrammatic).

We will now consider the influence of the type of gas on the ignition. The
fluorescent lamp usually contains argon, neon and mercury. It has been
found that ignition takes place more readily with mercury vapour present
than without mercury vapour. At room temperature the mercury-vapour
18 FLUORESCENT LAMPS

pressure inside the tube is about 0001 torr, whereas that of the argon
is several torr. The reason why this mixture ignites so well is that the argon
atom in the meta-stable state is capable of ionising the mercury atom
(Penning effect)*. This only occurs if the height of the meta-stable level of
the main gas (here the argon) is higher than the ionisation potential of the
admixture (the mercury). For the argon-mercury mixture these potentials
are 115 V and 104 V respectively.
The fact that argon-filled fluorescent lamps do not ignite quite so easily
at ambient temperatures below the normal is attributable to the fact that
the mercury-vapour pressure becomes so low that the above-mentioned
favourable condition no longer exists and the ignition voltage becomes
the same as for pure argon or argon-neon (see Fig. 8.32 and 8.35).

* F.M. Penning, NaU1rwiss. 15, 818 (1927); Zs. f. Phys., 46, 335 (1928).
Chapter 2

The discharge of the fluorescent lamp


W. Elenbaas

2.1 Why the low pressure mercury vapour discharge?


As pointed out in the previous chapter, ultra-violet radiation is produced
in the discharge and converted into visible light by the phosphor situated on
the inner wall of the tube. Since the sensitivity of phosphors is not the same
for all wavelengths - that is to say that the number of visible quanta we
obtain for 100 ultra-violet quanta absorbed by the phosphor depends on the
wavelength of the ultra-violet radiation (we call this number the quantum
efficiency in percent) - we will try to choose a discharge which produces as
much energy at the wavelength to which the phosphor is most sensitive.
Moreover we would like to use a phosphor which is very sensitive to ultra-
violet radiation, not too far away from the visible, because the energy dif-
ference between the ultra-violet quantum and the visible quantum is converted
into heat and therefore lost for light production. Thus with a quantum effi-
ciency of 100%, the energy efficiency equals Ae/Av if'Ae represents the wave-
length of the exciting quantum and Av that of the emitted visible quantum.
This ratio is closer to 1 the more Ae approaches the visible.
The possibility of conversion of a large part of the energy applied to the
discharge into emission of a single wavelength is only possible in the case of
20 FLUORESCENT LAMPS

the resonance lines. A gas or vapour must therefore be employed whose


resonance line lies in the ultra-violet part of the spectrum but not too far
from the visible region. In the case of a vapour, the vapour pressure at room
temperature should be of the order of w- 3 torr. If this vapour pressure is
lower, the temperature of the lamp must be increased in order to achieve
the desired vapour pressure necessary to produce resonance radiation at a
high efficiency. A high lamp temperature, however, involves heat losses (as is
the case, for instance, with low pressure sodium lamps). If all these require-
ments are to be met, the number of possibilities is very limited. The resonance
lines of rare gases occur far in the ultra-violet region, and the loss of energy
involved in the conversion into visible light is therefore high. Of the available
vapours, mercury is the only one that has a suitable vapour pressure, this
being 12 X I0- 3 torr at room temperature (20 aC). The resonance lines of
mercury lie at 185 and 254 nm, the latter being by far the stronger, so that
these are better for our purpose than the resonance lines of the rare gases.
Mercury vapour, therefore, does actually give the best results. As already
pointed out, the discharge must be one in which a large part of the applied
energy is transformed into resonance radiation. Now, the best conditions
under which the resonance lines occur are: low current density and low
vapour pressure. At low current densities there is not so much likelihood of
cumulative excitation of the atom, resulting in higher energy levels being
exited at the cost of the resonance levels. Low pressure is essential to avoid
appreciable absorption of the resonance radiation by the mercury vapour
itself. It is found that for the tube diameters used in practice optimum output
of the resonance radiation occurs at a vapour pressure of about 5 x 1o- 3 torr
(Fig. 2.2). At lower pressures the chance of collisions between electrons and
Hg atoms is too small, whereas at higher pressures too much of the resonance
radiation is reabsorbed. Rare gas (usually argon or A-Ne) is added to the
mercury vapour, to a pressure of a few torr, in order to increase the chances
of collision between the electrons and the Hg atoms. The pressure of the
rare gas must not be too high, as this increases the gas or volume losses
(see Section 1.4). The rare gas pressure is also determined by the starting
requirements (see Fig. I.13) and the required life (see Section 2.5.5). Since
the energy levels of A and Ne are higher than those of Hg, the rare gas
atoms are practically not excited.

2.2 Lamp dimensions


Let us suppose that a lamp is required which is to operate on a 220 V alter-
nating current mains and to consume W watts. The sum of the voltage drops
at cathode and anode is about 15 V, so that, in order to keep the electrode
losses relatively low, the arc voltage should be comparatively high, say
150 V. For reasons of stability and difficulty of ignition, however, it is not
usually practicable to go as high as that, a compromise being found in the
region of 100-125 V e.g. 115 V. The arc length multiplied by the voltage
gradient should thus give about 100 V (I 15 V minus 15 V drop at anode
and cathode). The current I is governed by the wattage and arc voltage
(apparent powerfactorofthe lamp about09-see Section 7.1). In our exam-
ple I= W/115 x09 A and the current density is I/(ind 2 ), d representing the
THE DISCHARGE OF THE FLUORESCENT LAMP 21
internal diameter of the tube. The requirement of a low current density suggests
the use of a large diameter tube. If, however, the tube is too wide, the
advantage of low current density is outweighed by the increase in self-absorp-
tion. Moreover, the temperature of the tube wall, and hence the mercury
vapour pressure, may become too low. This is due to the fact that, as a result
of the increasing tube diameter the gradient decreases, so that at a constant
current the amount of energy consumed per centimetre length of the tube
also decreases. At the same time the area of the tube wall increases in pro-
portion to the diameter. Both effects result in a lower wall temperature.
There is thus an optimum tube diameter for every value of the current and,
at the values of currents normally employed (!-It A), the appropriate dia-
meter lies between about 25 and 50 mm. At these diameters and at the low
pressures referred to above, the electrical gradient (V/cm) in the longitudinal
direction of the tube is low, e.g. of the order of 1-! Vfcm, so that, for a
100 V positive column voltage, the tube length has to be 1-2 m. For a high
efficiency we have thus to use long and wide tubes, with the consequent
disadvantages of awkwardness in handling and transport and the necessity
for large luminaires. On the other hand the surface brightness of these lamps
is so low that when placed in the direct field of vision they are considerably
less dazzling than filament lamps.
On 110 V supplies without transformers, the arc voltage has to be reduced
by about one half and the tube need be only half as long. In one sense this
is an advantage, but at the same time the efficiency of the shorter lamp is
lower, since the electrode losses are then relatively larger (see Section 2.5.4).

2.3 How much light can be expected from a fluorescent lamp?


To answer this question, let us first see how much light would be obtained
under idealised conditions, assuming that the mercury discharge transforms
all the applied power into radiation of 254 nm, that the quantum efficiency
of the fluorescent powder is I 00% (i.e. that a light quantum is produced for
every incident ultra-violet quantum), that the wavelength of this radiated
light quantum is that to which the human eye is the most sensitive (555 nm)
and finally that no absorption of it takes place in the powder itself. Naturally
such conditions could never be realised; nor, indeed would they be desirable,
as they would result in a light source emitting one wavelength only, such as
the low pressure sodium lamp. For certain purposes this might not be a
drawback, or might even be desirable, but for the purposes the fluorescent
lamp is normally used for, it would not do. It is, however, a simple matter
to calculate the efficiency of such an idealised light source. For every watt
of power consumed we then receive the equivalent of I W in rays of 254 nm
radiation, and this is in turn converted into rays of 555 nm. Since the energy of
a quantum of 555 nm represents only 254/555 of that of a quantum of254 nm,
we thus have 254/555 = 046 W emission of 555 nm. Now, 1 W of radiation
at 555 nm (maximum eye sensitivity) gives a luminous flux of 673 lm, so
046 W yields 310 lm. We thus obtain 310 lm/W, but it still remains to be
22 FLUORESCENT LAMPS

seen how far below this figure the practical result will be. The radiation spec-
trum of a fluorescent powder is never limited to a solitary spectral line, whilst,
moreover, in most fields of application a radiation is needed that will cover
the whole visible part of the spectrum, resulting in a visual sensitivity q over
the whole range of the spectrum from 0 to 1. The average value q will depend
on the energy distribution in the spectrum. The more the energy is concen-
trated around the 555 nm wavelength, the nearer q will approach 1.
The quantity q is defined as q = f E(A.) V(A.) dA.j f E(A.) dA. over the visible
part of the spectrum, where E(A.) and V(A.) represent the energy and the eye
sensitivity as a function of A.*. From the energy distribution of a fluorescent
powder, and the known eye sensitivity curve, the value of q may thus be
calculated.
In Table 2.1, q is given for a number of fluorescent powder mixtures
(characterised by the designation of the lamp-type), for the visible mercury
lines, and for the combination of the two. The approximate colour tem-
perature is given in the last column.
(The larger value of q for the visible Hg-lines for the lamps with the ab-
sorbing layer- colour 27, 32 and 37 - is caused by the absorption of the
blue lines; see Section 6.11).

Table 2.1

Philips q for the q for the approx.


colour
colour
fluorescent visible
q for the colour
designation combination
number powder Hg-lines temp. (K)

29 standard warm white 056 031 054 2 900


33 standard white 051 031 049 4100
54 standard daylight 043 031 042 6 500
32 de luxe warm white 040 050 041 2900
34 de luxe white 039 031 038 3 800
55 colour matching 042 032 041 6250
27 softone 036 052 037 2 700
37 trucolour 037 044 038 4100
57 daylight de luxe 034 031 033 7400
(artificial daylight)

The value of q for fluorescent lamps thus varies from about 035 (de luxe
lamps) to about 055 (standard warm white). In the following we will use
a q-value of 05.
The ratio of the energy of the light quantum to that of the 254 nm quan-
tum is rather better in the blue, but it is not so high in the red as the 046
derived above, so that this value can serve as an average for the 254 nm
radiation. For the 185 nm radiation the factor is 185/555 = 033, but,
since the 254 nm line is by far the stronger, we take a mean value of 045
for the energy ratio.

* If the fluorescent powder emits some radiation in wavelength regions where V(A.) is
zero we will extend the integration of E(A) over the whole range of emission.
THE DISCHARGE OF THE FLUORESCENT LAMP 23
A fluorescent powder having a quantum efficiency of 85% could be termed
very good, whilst a discharge that will transform two thirds of the applied
power into resonance lines is the best that can be achieved (the tube must
then be chosen so long that electrode losses become relatively small, in con-
junction with a diameter and current that will give optimum results). The
luminous flux produced by 1 W of power is then 673 x045 x050 x085 x2/3
R:J 85 lm, this giving a theoretical efficiency of 85 lmjW. This is indeed
about the efficiency obtained nowadays for long tubes (we have neg-
lected the electrode losses) and standard white (ij R:J 05) under favourable
conditions. Needless to say, it is easier to achieve higher efficiency values the
more the emission band of the phosphor is concentrated around the 555 nm
line, since the ij-value is then larger; the colour rendering will not be so good,
however, but for certain purposes this will not matter.

2.4 The energy balance of the fluorescent lamp


Fig. 2.1 is a diagrammatic representation of the conversion of power in a
fluorescent lamp. It is assumed that 100 W of electrical power is consumed,
and the diagram shows the different forms of energy into which this amount
of power is converted, the width of the bands being roughly proportional
to the quantity of energy flowing through them. The energy values indicated
are of course merely averages, since they vary from one lamp type to an-
other, dependent on current density, mercury vapour pressure, fluorescent
powder and so on. Of the 100 W input, about 2 W are converted directly
into visible light by the discharge, about 60 W into ultra-violet radiation of
which the larger part is emitted in radiation of 254 nm and 185 nm; the
remainder is dissipated as heat in the discharge, at the walls and at the
electrodes.
The energy of 60 W of 254 nm and 185 nm radiation reaches the fluores-
cent powder. On the basis of an average energy factor of 045 for emitted
to absorbed quanta, together with a quantum efficiency of 85 %, 60 x 045 x
085 = 23 Ware converted into visible light, so that in total25 W of visible
radiation are emitted. Taking for ij the value 05, 25 x 673 X 050 = 8 500 lm
is obtained.
The wall of the tube has to dissipate 75 W. Under conditions of equilibrium
about 60% of this is removed by conduction and convection and 40% by
radiation, but this ratio depends largely on the temperature of the tube itself
and therefore also on the tube dimensions. The method of suspending the
lamp- either naked or enclosed in a luminaire- also affects the distribution
of the energy loss by the tube wall over radiation and conduction plus con-
vection.
Fig. 2.1 demonstrates clearly why the fluorescent lamp is so satisfactory
from the point of view of the radiation of heat; of the 100 W consumed by
the lamp, about 25 + 30 = 55 W takes the form of radiation. The lumi-
nous flux is 8 500 lm, i.e. approximately 0006 5 radiated watt per lumen. The
tungsten filament lamp presents quite another picture in that a 100 W lamp,
in which some 90 W is converted into radiation, yields less than 1 500 lm,
24 FLUORESCENT LAMPS

TOOW

Convers;on
in the phos-
phor

..- Conversion at
the wall

8500/m

Fig. 2.1 Diagram representing schematically the conversion of energy in a fluorescent


lamp. Of the 100 W consumed, roughly 2 W are directly converted to visible light and
60 W to other radiations mainly of 254 and 185 nm. The rest is lost: 15 W electrode losses,
23 W volume and wall losses of the discharge; these volume and wall losses are manifested
at the tube wall as heat. The 60 W of ultra-violet radiation reaching the powder yields
about 23 W in visible light and some 37 Win heat, the latter again dissipated at the wall.
Of the 75 W arriving at the tube wall about 30 W are radiated (infra-red) and 45 W are
disposed of by thermal conduction and convection to the surroundings. The 25 W of
visible radiation give approx. 8 500 lm (dependent on the value of q, thus on the energy
distribution of the light).

which means that, per lumen, the radiation amounts to about 0.06 W or
nearly 10 times as much as in the fluorescent lamp. For the same illumina-
tion, then, the tungsten lamp gives about 10 times as much radiant energy
as the fluorescent lamp, resulting from the high efficiency of the latter type
of lamp, compared with the filament type, as well as from the lower total
radiation of the fluorescent lamp.
THE DISCHARGE OF THE FLUORESCENT LAMP 25

2.5 The efficiency as a function of different parameters


The luminous flux emitted by a fluorescent lamp is dependent on the
quality and layer thickness of the fluorescent powder, as well as on the
amount of incident radiation of the wavelengths of 254 and 185 nm. The
lamp efficiency therefore also depends on these factors. As far as the quality
of the powder is concerned, the reader is referred to Chapter 4. The fact that
the layer thickness is important will be readily appreciated when it is remem-
bered that there can be no fluorescence in the absence of a fluorescent powder,
and that in the case of a layer of infinite thickness, the light cannot emerge
from tbe lamp because it is produced at the inside of the coating. There is
thus an optimum layer thickness which guarantees absorption of almost all
the 254 nm radiation, with a minimum absorption of the fluorescent light.
Let us now consider the third factor, the amount of radiation at 254 and
185 nm received by the powder for a given amount of power. The intensity
of these lines may depend on the mercury vapour pressure, the current, the
tube diameter, the arc length, the gas filling, the gas pressure and the form
of the current. The quality of the electrodes is also important, especially the
work-function of the cathode and the surface area of the anode; the smaller
the voltage drop at anode and cathode, the higher the efficiency. Let us now
deal in turn with all these parameters from the aspect of the influence upon
the intensity of the resonance radiation.

2.5.1 The mercury vapour pressure


Varying the mercury vapour pressure and holding all other parameters
constant, it can be anticipated that an optimum pressure will be found. If the
mercury vapour pressure is too low, there is insufficient likelihood of colli-
sions between electrons and Hg atoms, whereas excessive vapour pressure
results in too much absorption of the resonance radiation. Somewhere be-
tween the two extremes, therefore, the most satisfactorypressuremustoccur.
Experimentally, the performance as a function of the mercury vapour pres-
sure can best be determined by operating the fluorescent lamp in a water-
bath of which the temperature is variable, and measuring the amount of light
emitted at a constant current. The consumption of power at constant current
will vary slightly, since the gradient depends slightly on the vapour pressure.
The relative variation in efficiency 1J measured in this way is plotted in Fig.
2.2, in which the vapour pressure of the upper scale corresponds with the
temperature of the water-bath of the bottom scale; the vapour pressure is
thus the saturated vapour pressure of mercury at the corresponding tem-
perature. Maximum efficiency is obtained between 32 and 47 ac, corres-
ponding with a vapour pressure of 3-lOx w- 3 torr.
With a load of 40 W in a tube of 4 ft length and 30-40 mm diameter
operated in a surrounding of 20-25 ac the wall temperature acquires a tem-
perature of approximately 40 C, which gives the optimum mercury vapour
ressure. In some cases, however, a larger load is wanted in the same lamp.
26 FLUORESCENT LAMPS

-
-P
05 2 3 4 5 7.5 10 15 20 30 40 50 KI0- 3 torr
100 -L_ ~I
/
!.--"'
1'--
8 /

f
J
[7
60
f7
J
40
v
20
i
0
0 10 30 60

Fig. 2.2. Relative efficiency 1} of a 40 W fluorescent lamp, as a function of the mercury


vapour pressure. The pressure is indicated at the top of the diagram, and the temperature
at which this pressure occurs on the bottom scale.

If we then do not take special measures, the mercury vapour pressure will
become too high and the efficiency will be lower for two reasons : the larger
current density and also the excessive mercury vapour pressure. We may
correct the latter in several ways, so that only the higher current density
then causes a lower efficiency (in Fig. 2.7 'f/A-+'f/ 8 by adjusting the op-
timum vapour pressure).

Fig. 2.3. Cross-section of the ' Power-


Groove' lamp.

Fig. 2.4. 'Very high Output lamp'. The coldest spot which determines the mercury
vapour pressure is situated behind the electrode.
THE DISCHARGE OF THE FLUORESCENT LAMP 27

==t$~------------~vr-----------~@d==~
Fig. 2.5. Double flux lamp. The mercury vapour pressure is determined by the tempera-
ture of the side tube where all the liquid mercury condenses.

Fig. 2.6. Instead of using a side


tube, the lower temperature may
also be created by an extra
cooling of a part of the tube wall.

The mercury vapour pressure may be lowered in the following ways:


1. By increasing the cross-section of the discharge tube, preferably combined
with a non-circular cross-section 1 >, since this creates cold spots. Moreover
a cross-section as indicated in Fig. 2.3 has the advantage that the distance
an ultra-violet quantum has to travel before striking the fluorescent pow-
der is on an average shorter than with a circular cross-section.
2. By creating a cold spot in the tube either by increasing the distance be-
tween at least one of the electrodes and the end of the tube as shown in
Fig. 2.4 2 >or by applying a small side tube as indicated in Fig. 2.5 3 >, or by
cooling a spot of the tube from the outside by pressing a good heat con-
ducting pin against it (Fig. 2.6).
3. By keeping the temperature high, but applying an amalgam instead of
pure mercury, so that notwithstanding the high temperature, the mer-
cury vapour pressure attains the right value. In practice a Hg-In amalgam
is used 4 >. By variation of the ratio of the two elements, the Hg vapour
pressure may be varied at a given temperature. The application of an
amalgam is a very elegant method, but has two drawbacks:
(a) At room temperature the mercury vapour pressure is also lower than
with pure mercury. As regards starting, the lamp thus behaves as a
normal lamp at a lower temperature. The lowest temperature at which
the lamp starts reliably is thus higher. Amalgam should therefore not
be used in lamps which have to start at low temperatures unless
measures for reliable ignition are taken.
(b) Because the mercury vapour pressure at starting is lower, the time
necessary to reach say 90% of the ultimate light output, is longer. It
is possible, however, to get rid of this drawback by also applying
Indium on the ring situated around the electrode coil (see Fig. 5.8).
This ring is heated up rapidly after ignition of the lamp and the
amalgam of the ring therefore also gives off mercury rapidly, so that
the increase of light output after ignition is accelerated considerably.
We just considered lamps in which the input was increased above the
normal value at which the optimum efficiency occurs. It is also possible that
the load of the lamp is normal, but that the lamp is operated in a hot sur-

ll J. 0. Aicher and E. Lemmers, Ill Eng. 52, p 579, 1957.


2 > F. J. Waymouth, W. Calvin Gungle and Ch. W. Jerome, The Sylvania technologist 9,
p 102.
3 > H. J. J. van Boort and D. Kolkman, Philips techn. Rev. 19, p 333, 1958.
4 > K. Eckhardt and B. Kiihl, Lichttechnik 22, p 389. 1970.
28 FLUORESCENT LAMPS

roundings (for instance a small badly ventilated luminaire with several lamps
in it). The measures treated above to keep the vapour pressure sufficiently
low, will also then apply and the solution under 3. is very effective, because
it works when the whole tube is situated in hot surroundings. The solutions
according to 1. and 2. will fail when the ambient temperature is higher than
about 40 oc, except the solution of Fig. 2.6 when the heat conducting pin
protrudes into colder areas so that it can carry heat away from the tube wall.
It is also possible that the mercury vapour pressure is too low because the
current is small or the surrounding temperature is low. In that case the lamp
may be operated in an outer glass tube in order to increase the temperature
of the discharge tube.

2.5.2 The current


All other factors being constant, it is to be expected that an increase in
current will be accompanied by a drop in efficiency of the resonance radia-
tion, since higher current density increases the risk of cumulative excitation.
Measurement, which again may be carried out in a water-bath in order to
keep the mercury vapour pressure constant, confirms this supposition. In
Fig. 2. 7 the full line represents the efficiency 'YJ at constant mercury vapour
pressure.

p
too
'l / 1 - """'=::
t::::- ~
r----... --
I ~::- -.....;.;;: .......
80 ..... ....
Fig. 2.7. ftz I
- - Efficiency as a function of the I
current at optimum mercury vapour 60
pressure (lamp operated in a water-bath I
at the optimum temperature).
- - - - - Efficiency as a function of the 40 I
current at mercury vapour pressures as I
I
determined by the current (at a given
temperature of the surrounding air). 20 i
The lamp attains optimum pressure at I
the current I 1 . By reducing the vapour
pressure to the optimum one at current
12 the efficiency increases from 'YJA to 0 i
0 I, fA
'YJB ~I

Let us suppose that this curve relates to the optimum mercury vapour pres-
sure. If the lamp is now operated in air, the current can be adjusted to pro-
duce this optimum pressure. Let the value of the current at which this occurs
be I 1 resulting in an efficiency 'Y} 1 (point P); when the current is varied, the
mercury vapour pressure occuring at I> I 1 will be higher, and in the case
of I < I 1 lower than the optimum value. Thus if the lamp is operated in
air at different currents the efficiency will follow the broken line in the dia-
gram, which touches the full line at P. At very low current values the pres-
THE DISCHARGE OF THE FLUORESCENT LAMP 29
sure drops to such an extent that the efficiency falls in spite of the lower
current density. To increase the efficiency in this case we may apply a thermal
heat insulation (see also Section 2.5.1). At the current / 2 we can increase
the efficiency from 'YJA to 'Y/B by decreasing the vapour pressure to the op-
timum pressure by one of the methods given in Section 2.5.1.

2.5.3 The tube diameter


At constant current and mercury vapour pressure, there will be an op-
timum tube diameter because with a very small diameter the current density
will be too high (resulting in more excitation of the higher energy levels),
whereas with a very large diameter the absorption of the 254 nm radiation
will be high because the ultra-violet quanta travel a long average distance
before reaching the fluorescent powder. Moreover, under conditions com-
parable with actual practice, in which the vapour pressure is also dependent
on the diameter of the tube, a larger diameter may result in too low a vapour
pressure. Too high a mercury vapour pressure at high current densities may
be prevented by creating a cold spot somewhere in the tube (see Section 2.5.1).
With increasing current the optimum tube diameter will increase.

2.5.4 Tube length and the combined voltage drop at anode and cathode
These two factors affect the efficiency because that part of the arc voltage
which is of practical utility is dependent upon them. In effect, the arc voltage
is equal to the sum of the voltage drop at anode and cathode, plus the voltage
of the positive column, but only the latter is of importance for the produc-
tion of light. The energy consumed in the voltage drop at anode and cathode
is lost; since the anode and cathode fall are independent of the length of the
positive column, whereas the voltage over the column is proportional to the
length, the efficiency increases with increasing arc length. This relationship
can in fact be easily calculated. Let the total lamp voltage be Vi and the
anode and cathode fall Va+c; all other conditions being equal then, the
efficiency is proportional to (Vi- Va+c)/Vi. Fig. 2.8 illustrates the effi-
ciency as a function of the arc length in respect of a gradient E of075 Vjcm
(V1 = Va+c + 075) and Va+c = 15 V (curve 1), Va+c = 25 V (curve 2) and

Fig. 2.8. Efficiency 'YJ as a m.----.-----.-----,------,


function of the length of the
column. The efficiency in the
absence of any electrode
t
'L

losses is taken to be I 00.


1. E =! V/cm; anode + 50t---F-7"'--+-::"ii"""""---+- -::::;;;;o-~---~
cathode fall = 15 V
2. E = t V/cm; anode +
cathode fall = 25 V
3. E = t V/cm; anode +
cathode fall = 100 V
4. E = It V/cm; anode + 50 tOO
cathode fall = 100 V 150 200
-l{cm)
30 FLUORESCENT LAMPS

Va+c = 100 V (curve 3). The efficiency when I= oo is placed at 100 (when
I= oo, the efficiency is unaffected by Va+c) Curves I and 2 relate to thermal
emission of the cathode (and conditions being such that the gradient is
075 V/cm), whereas curve 3 refers to a cold-cathode tube; in this case a high
efficiency is attained only when I is very large, and this is the reason why
cold-cathode lamps are usually of considerable length. Such lamps are
generally operated on a lower current (e.g. 100 rnA) through a narrow tube
(e.g. 25 mm diam.), resulting in a high electrical gradient; for equal lengths
therefore, the effective part of the voltage increases with decreasing tube
diameter. Curve 4 in Fig. 2.8 relates to a gradient of 15 V/cm with Va+c =
100 V; in comparison with curve 3, the same efficiency is obtained for a tube
of only half the length.

2.5.5 The rare gases

EFFECT OF PRESSURE
Here again, there is an optimum value, for, without the rare gas and with
only mercury vapour in the tube, the unimpeded path of the electron is very
large (exceeding the diameter of the tube) and the electron is more likely to
fly at high speed against the wall or the electrodes than to collide with and
excite a mercury atom. When the gas pressure is very low, therefore, losses
at the wall of the tube are excessive. On the other hand, we have seen that
when the gas pressure is high the number of collisions between electrons and
gas atoms is so large that, although each collision results in the transfer of
only a small amount of energy, the total amount of energy thus converted
into heat is considerable (loss by elastic collisions - see Section 1.4). Thus
an optimum gas pressure exists that will ensure a minimum value of the sum
of losses caused by elastic collisions and losses at the wall. This pressure will
moreover depend on the tube diameter, since the latter governs the losses
at the wall. For larger tube diameters, the optimum will occur at a larger
mean free path, thus at a lower gas pressure. These optimum gas pressures
are mostly so low that in practice lamps are filled with higher pressures in
order to achieve a sufficient lamp-life, the latter increasing rapidly with
increasing gas pressure. For the influence of the pressure on the ignition see
Fig. 1.13.

EFFECT OF THE KIND OF GAS USED


The transfer of energy that takes place in a collision of the elastic kind
between an electron and an atom is smaller the heavier the atom concerned.
For a given current, vapour pressure, tube diameter and rare gas pressure,
the efficiency of the positive column of a fluorescent lamp, as a function of
the rare gas used, therefore increases in the following order: He, Ne, A, Kr,
Xe. The most widely used gas is argon, for very simple reasons. He and Ne
usually give lower lamp efficiencies and moreover a shorter life (owing to a
greater voltage drop at the cathode) Besides they are also more expensive
THE DISCHARGE OF THE FLUORESCENT LAMP 31
gases. This leaves,besides A, only Kr and Xe, and, as a matter of fact,
krypton has been employed in these lamps. Xenon is not so plentiful and,
since it is therefore much more costly, its use is precluded. In many cases,
the advantage to be gained by using krypton instead of argon is only very
slight; moreover, starting does not take place so readily in this gas and the arc
voltage is lower than in the case of argon. For a lower arc voltage the ob-
vious step would be to employ a lower mains voltage or, keeping the latter
unchanged, to use a longer tube. Both measures, however, result in more dif-
ficult lamp starting. For these reasons the use of krypton for fluorescent
lamps is very limited. An addition of Ne to the A is often used in order to
increase the arc voltage of the lamp at a given current, resulting in a higher
energy input and a higher lumen output. Additions of neon up to about
30% do not affect the efficiency.

2.5.6 Effect of the form of the current


Since the efficiency of production of resonance radiation depends on the
cunent density (see Section 2.5.2) it must be expected that the current form
has an influence on the lamp efficiency. Since the efficiency decreases at
higher current densities, we expect the highest efficiency with direct current
and the lowest efficiency when the current is peaky, causing a high momen-
tary current density. This is in agreement with experiments. The difference
in efficiency between d.c. and sinusoidal a.c. is only small, whilst a square
wave form is equivalent to d.c. Peaked currents, as obtained with only a
capacitor in series with the discharge, result in a low efficiency. Increasing
the frequency in this case, however, again gives a more sinusoidal current,
so that at high frequencies stabilisation with only a capacitor is a good
proposition *.
Although d.c. gives the highest lamp-efficiency, it is in practice not a nice
solution for two reasons:
1. The d.c. discharge has to be stabilised with a resistor, the energy dissipated
in it is either lost, or if it is made in the form of an incandescent lamp,
gives only a low efficiency, so that the overall efficiency is lower than for
an a.c. operated fluorescent lamp stabilised with a choke.
2. With d.c. operation the mercury ions move, on an average, towards the
cathode end of the tube, so that after operating for some time, that part
of the mercury which was at the anode end of the tube is evaporated and
transported to the cathode end. As a result the emission of 254 nm radia-
tion at the anode end of the tube decreases for lack of mercury atoms,
and that part of the tube emits less light. The polarity of lamps operated
on direct current has therefore to be reversed periodically, the actual
interval depending on various factors (in practice some 5 h). See also
Section 8.9.

* J. H. Campbell, H. E. Schultz and D. D. Kershaw Ill. Eng. 48, 95 (1953).


Chapter 3

Luminescence, fluorescence and phosphorescence


J. L. Ouweltjes

3.1 Incandescence and luminescence


In Section 1.3 we have seen that light is emitted as the result of an electronic
transition from a level of higher energy towards a level of lower energy. This
applies quite generally for solid materials such as the tungsten coil of an
incandescent lamp as well as the luminescent materials used in fluorescent
lamps. At room temperature the electrons occupy the lowest possjble energy
levels. Therefore, if we want to produce light, means have to be found to
bring atoms into the excited state. In the older forms of light sources, this is
effected by raising the material to a high temperature. In the candle or the
paraffin-oil flame, extremely fine particules of carbon are produced and these
are raised to incandescence by the heat of the gases also formed by combus-
tion. The light of the incandescent gas mantle is likewise the result of a
transfer of heat, in this case produced by the combustion of the gas. Finally,
in the electric incandescent lamp, the filament is heated by the electric cur-
rent. All these light sources, therefore, have as a common feature that the
light producing material is at a very high temperature. This method of pro-
ducing light is called incandescence. There are many other ways, however,
of making materials emit light. As discussed in Chapter 1, in a gaseous dis-
charge the energy required to excite the atoms may be produced by collisions
LUMINESCENCE, FLUORESCENCE AND PHOSPHORESCENCE 33
between free electrons and the atoms of the gas. The familiar red neon sign
used for advertising purposes is an example of a light source based on this
principle. Another means of exciting materials is irradiation with electro-
magnetic radiation of a suitable wavelength. It has also been explained in
Chapter 1 that this is what actually happens in a fluorescent lamp, the ultra-
violet generated in the gas discharge excites the fluorescent material on the
wall of the tube. Still another way of excitation, familiar to everyone nowa-
days, is the bombarding of the material with cathode rays, as is done in the
picture tube of a television set. The picture is produced by a cathode-ray
beam of variable intensity, scanning the front face of the picture tube, which
has been coated with a fluorescent screen. Technically unimportant methods
are occasional excitation by mechanical means such as crushing and the
generation of light during chemical reactions. A well-known example of the
second kind is the glow of elementary phosphorus in the dark. The phos-
phorus undergoes a slow oxidation which is accompanied by emission of
light. This phenomenon is fundamentally different from what happens in the
fluorescent lamp or in the TV picture tube, where the light emitting material
undergoes no chemical change. All the phenomena in which the emission of
radiation by a body is not due to incandescence are classified under the term
luminescence.
Materials showing luminescence are often callerl phosphors. According to
the type of excitation, distinction is made between photoluminescence (exci-
tation by electromagnetic rays), cathodoluminescence (excitation by cathode
rays), chemiluminescence, etc.
In general, incandescence is a property that is primarily determined by the
temperature of the radiator. The chemical composition of the material is of
only secondary importance. A body having a temperature of I 500 oc has
very much the same appearance whether made from a metal or from a
refractory oxide. On the other hand, luminescence is strongly restricted to
well defined chemical compositions. Luminescent materials therefore must
be carefully synthesised, otherwise the excitation energy will be dissipated
in the form of heat, and no visible emission will occur.

3.2 Some practical aspects of the luminescence of solids


Before discussing the theoretical basis of the luminescence of solids we
will first describe some of the most important experimental aspects.
One of the most remarkable facts in the luminescence of solids is that the
energy distribution of the emission is in general independent of the excitation
mechanism.
Willemite, the phosphor used in combination with a mercury discharge
in green sky signs, always shows the same emission curve, whether excited
by 254 nm radiation, by the neon line 73 nm, or by cathode rays. The same
holds for calcium tungstate, which may be excited either by 254 nm radiation
or by X-rays, and for zinc sulphide phosphors, that show the same lumines-
cence under either ultra-violet radiation of all wavelengths or cathode rays
and X-rays.
34 FLUORESCENT LAMPS

Though the spectral characteristics of the emission are independent of the


manner of excitation, this does not apply to the efficiency of the luminescent
process. In particular, for excitation by ultra-violet the efficiency depends on
the wavelength. The total efficiency is the result of two factors: the absorp-
tion of the incident ultra-violet and the efficiency with which the absorbed
ultra-violet is converted into visible radiation. Most phosphors used in
fluorescent lamps show an absorption that is negligible for wavelengths
greater than 300 nm. With decreasing wavelength, absorption increases and
as a consequence the luminescence gets stronger. However, not all of the
absorbed radiation need necessarily lead to luminescence. Variations in the
intensity of luminescence occur which cannot be ascribed to a difference in
absorption of the ultra-violet, but that must be connected with the efficiency
with which the absorbed energy is converted into visible energy. This efficien-
cy also depends on the wavelength of the radiation which is used for excita-
tion. We must therefore distinguish between the absorption spectrum and
the excitation spectrum. For some phosphors they are very similar, for others
they are quite different. An example is given in Fig. 3.1, taken from a paper
by Blasse and Bril *,where two terbium activated phosphors are compared.

~rov~'oor---~~~~--~----,
Tb3 +quantum Reflection(%)
output 80

i 60
i
40

20

280 260 240 220


--Aoxcfnm) --Aoxcfnm)

Fig. 3.1. Relative-excitation spectra of the terbium fluorescence and diffuse-reflection


spectra of yttrium tungstate: Tb and yttrium titanate-tantalate: Tb. After Blasse and Bril.
Instead of the absorption the reflectance ( = 1 -absorption) has been plotted.

Many substances continue to emit light after excitation has been stopped.
The theoretical background is that the excited atoms gradually return to the
ground state. In consequence, the intensity of the light gradually decreases
as more and more of the atoms have returned to the ground state. The time
in which this happens (the decay time) may vary from w-s s to several
hours. For materials having such a long afterglow the return to the ground
state is indirect.

* G. B!asse and A. Bril, Philips Research Rep., 22, 481 (1967).


LUMINESCENCE, FLUORESCENCE AND PHOSPHORESCENCE 35
The phosphors used in TV tubes must have a negligible afterglow, other-
wise the picture will be blurred. For fluorescent lamps an afterglow of about
001 s is desirable in connection with the dark period in the lamp cycle,
which is of the order of 001 s.
The emission of light simultaneously with the excitation is usually referred
to as fluorescence, whereas a noticeable afterglow is designated by the term
phosphorescence. A sharp distinction cannot be made, however, for strictly
speaking the absorption and emission in one atom never occur simultane-
ously, so there is always some delay between them. The typical phosphores-
cence depends strongly on temperature. At higher temperatures the light is
released more rapidly than at a lower temperature. When such a phosphor
is excited at liquid air temperature and then gradually warmed up, the inten-
sity of the emitted light as a function of time (the so-called glow curve)
provides interesting information on the nature of the excited state. This will
be discussed in more detail in Section 3.5.
Fluorescence generally persists at lower temperatures, but at higher tem-
peratures quenching occurs. Sometimes this quenching begins below room
temperature, and these phosphors are unsuitable for practical purposes.
Other phosphors retain a high efficiency up till 300 ac: when of a suitable
colour they may be used to improve the spectral characteristics of the high
pressure mercury vapour lamps.

3.3 The emission and absorption spectra ofsolid substances


The example of the mercury atom discussed in Chapter 1 clearly illustrates
that even in the individual atoms and ions, such as they exist in the gaseous
state, there is a large number of energy levels between which optical transi-
tions of the electron can occur and as a consequence the emission and ab-
sorption spectra show many spectral lines.
The complexity of the emission and absorption spectra increases, however,
when we turn to more complicated atomic structures, such as molecules, or
to materials in the condensed state, such as liquids and solids. The principle
according to which the difference between two energy levels determines the
wavelength of the radiation is still maintained, but the energy is no longer
exclusively dependent on the configuration of the electrons. For instance, in
a diatomic molecule the total energy depends on the rotation of the molecule
and also on the mutual vibrations of the atoms. Like the electronic energies
the rotation and vibration energy are quantised. The distances between the
various rotation and vibration levels are much smaller, however. As these
rotational and vibrational levels are superimposed upon the electronic levels,
the latter will split up into broad energy bands. This leads to broad absorp-
tion and emission spectra.
In the solid state rotation of groups of atoms occurs only in exceptional
cases, but the vibrations of the atom in the crystal suffice to cause broad
absorption and emission bands in this case also.
The characteristics of the absorption spectrum are important in connection
with the application of phosphors. Most phosphors consist of a host crystal
36 FLUORESCENT LAMPS

which does not luminesce by itself, with a small amount of a foreign consti-
tuent that causes the luminescence, and is therefore called the activator.
In the gaseous state, when we are dealing with the individual atoms, we can
distinguish between transitions towards levels of higher energy, but still
belonging to the atom, and transitions which lead to ionisation. In the same
way, in a phosphor consisting of a host crystal and an activator we can in
principle expect also two kinds of transitions, those in which the electron
remains localised in the atom to which it belongs, and those in which it is
set free and renders the material conducting (photoconductivity). As this
applies to the activator as well as to the host crystal, this provides us with
four possibilities. Often, but not necessarily, the absorptions due to the
activator require less energy and therefore will be found at longer wave-
lengths. In general, the activator concentration is small, of the order of per-
cents or even less, which implies that the absorption is not as strong as that
of the host lattice. We thus get an absorption spectrum of the form repre-
sented in Fig. 3.2. For most lamp phosphors absorption of the 254 nm reso-
nance line is due to the activator, and the excited electron remains localised
in the activator or its immediate surroundings.

Abs

f Lattice
Activator
absorption
absorption I
I
I
I

,'t"
II I \
I \

\ ,,
.... Fig. 3.2. Absorption spectrum
-A. of a luminescent material.

We now come to the relative positions of the absorption and emission


bands. In the gaseous state, especially when the pressure is very low, excited
atoms will return to the ground state without interaction with the surroun-
dings (resonance radiation). In a solid material this is very unlikely to hap-
pen, even the short life-time of an excited state ( ~ w- 8 s) is sufficient to
enable a complete exchange of energy with the surroundings. The atom may
pick up some additional energy from the vibrations, or, and this is the more
general case, it loses energy. In non-luminescent substances the electronic
energy of the excited state is converted completely into vibrational energy,
which means into heat. In phosphors the electron returns to the ground state
under emission of radiation, but the wavelength is nearly always different
from that used in excitation. In the more general case, that of a loss of energy
to the surroundings, the emission is at longer wavelengths than the absorp-
tion. This rule is known as Stokes' law. It amounts to a complete separation
of absorption and emission spectrum. This applies also for most of the lamp
phosphors, for which absorption starts at about 300 nm, whereas the emis-
LUMINESCENCE, FLUORESCENCE AND PHOSPHORESCENCE 37
sion is in the visible spectrum. Occasionally, however, we meet phosphors
for which the emission and the absorption spectrum partly overlap (Fig. 3.3).
Excitation by means of monochromatic radiation of wavelength a will give
rise to the complete emission spectrum. For that part which is right of a,
Stokes' law applies, but for the radiations with wavelength smaller than a,
Stokes' law is violated. In this case the emission spectrum depends on the
activator concentration. The phosphor absorbs the short wavelength tail of
its own emission, and as this absorption will increase with increasing activa-
tor concentration, the net effect will be that the short wavelengths in the
emission are suppressed as the activator concentration increases.

Fluorescence
intensity

Fig. 3.3. Stokes' law. Partly


overlapping absorption and
emission curves.

3.4 Some further considerations on the electron transitions in-


volved in luminescence.
A comprehensive description of the luminescence of solids should in the
first place entail the knowledge of the actual electronic transitions involved
in the absorption of the ultra-violet and the emission of visible light. As
already mentioned, in many instances the occurrence of luminescence in a
substance depends on the presence of a small amount of a foreign constituent,
known as the activator. In these cases the energy levels involved in the elec-
tron transition should be connected in some way or another with the activator
atom. In the simplest case the absorption of the ultra-violet radiation promo-
tes an electron from a level of lower energy to a level of higher energy, fol-
lowed by a return to the original level. The host crystal plays no essential
part; it can primarily be considered as a diluent that keeps the activator
atoms separate. This is an important point in order to get an efficient energy
conversion, for if the concentration of the activator atoms gets too high this
generally results in a drop in the emission intensity. The influence of the host
crystal cannot be completely neglected, however. Especially when the elec-
tron jump occurs in the outer electron shells the position of both energy
38 FLUORESCENT LAMPS

levels is affected, so that the position of both the absorption spectrum and
the emission depends on the host crystal.
This means that the direct environment of the activator atom is the first
thing to be considered if an attempt is made to connect its energy levels with
those of the free atom. For a long time it has been thought that there should
be something special about the position of the activator in the crystal lattice;
we know now that the activator atom simply replaces one of the normal
atoms in the host crystal. In zinc silicate activated by manganese the manga-
nese atoms are on crystallographic positions normally occupied by zinc
atoms, in potassium chloride activated by thallium the thallium atoms substi-
tute for the potassium atoms, etc.
In the last case a successful attempt has been made to calculate the posi-
tions of the various energy levels (and thus of the absorption and the emis-
sion spectra) starting from those in the free thallium ion and correcting for
the influence of the surrounding ions. The potassium chloride lattice is a
simple one, it consists of a three-dimensional cubic array of alternating
potassium and chlorine ions (Fig. 3.4), but even then the calculations are
extremely difficult. For the more complicated crystal lattices calculations
are out of the question.

Fig. 3.4. Crystal lattice of potassium


chloride. potassium ions; 0 chlorine
ions and 0 a foreign positive ion repla-
cing a potassium ion.

But apart from this qualitative simple mechanism of luminescence there


are many cases in which the phenomena are more complex. When the exci-
ting radiation is absorbed not by the activator, but by the atoms of the host
lattice, then the excitation energy must ultimately be transferred to the acti-
vator, which often amounts to a transport through the host crystal to one of
the next neighbour atoms of the activator, followed by a transfer to the acti-
vator. This transport of energy can also occur by different mechanisms, one
of which will be described more elaborately. It occurs in the sulphide phos-
phors that are universally used in black and white television picture tubes.
The usual phosphor mixture consists of a blue emitting zinc sulphide and a
yellow emitting zinc cadmium sulphide. In both cases silver is the activator.
In the television picture tubes they are excited to luminescence by means of
cathode rays, but they also fluoresce quite well when excited by 254 nm
radiation, which is in the region of the host crystal absorption. The substi-
tution of zinc by silver in zinc sulphide is accompanied by the incorporation
LUMINESCENCE, FLUORESCENCE AND PHOSPHORESCENCE 39
of an equivalent amount of a so-called coactivator. Chlorine is the most
common coactivator. It replaces a sulphur ion, and as it has one electron
more than the sulphur ion, it donates an electron to the silver ion, which
has a deficit of an electron as compared with the zinc. In this way two local
energy levels are formed, an occupied one at the silver ion (Din Fig. 3.5) and
an unoccupied one at the chlorine ion (Gin Fig. 3.5). The position of these
energy levels is between those of the outer electrons of the sulphur ions (the
valence band, s-- in Fig. 3.5) and those of the electrons which can move
freely throughout the crystal lattice (the conduction band, Zn+ in Fig. 3.5).

Fig. 3.5. Energy levels of the electrons


in a zinc sulphide phosphor.

The electronic transitions involved in the luminescence of such a zinc


sulphide phosphor when excited by 254 nm radiation may be described as
follows. Absorption of a 254 nm quantum causes an electron to jump from
the valence band into the conduction band (A ~ B in Fig. 3.5). Like the
electron in the conduction band, the vacant electron site (electron hole) left
in the valence band can move freely from one sulphur ion to another. When
it comes in the vicinity of an activator ion (A~ C) the hole will be filled
with the electron occupying the level D, and in this way the electron hole
becomes trapped. If the electron moving the conduction band ultimately
reaches this electron hole (B ~ E) they will recombine, which is accompa-
nied by the emission of light (E ~D).
Phosphorescence is connected with energy levels which, like G, are nor-
mally unoccupied and are lying slightly below the conduction band. These
energy levels function as traps for the electrons in the conduction band, and
it only depends on their position and the temperature of the phosphor
whether a trapped electron will soon return into the conduction band, or
whether it will remain trapped for a long time. The lower the temperature,
the longer the electron will remain in the trapping level G. The introduction
of trapping levels is not necessarily restricted to the coactivator atoms, other
foreign constituents may have the same effect. It will be clear that each type
of foreign ion create<; its own trapping level. The closer the trapping level is
situated to the conduction band, the easier it will be emptied. In the glow
curve (see p. 35) this is reflected in the temperature at which the intensity
shows a maximum. The closer the trapping level lies to the conduction band
the lower will be the temperature at which the peak intensity is observed.
In this way the study of the glow curve may help to elucidate the cause of
the afterglow.
40 FLUORESCENT LAMPS

Transport of excitation energy also occurs in the so-called sensitised phos-


phors. They contain two activators. One of them acts also as an activator
without the other being present, whereas the emission of the other only
occurs when both are present. To give an example, calcium silicate (CaSi03)
to which a small amount of manganese has been added will not show any
fluorescence when irradiated with 254 nm radiation. But when it also con-
tains a small amount of lead, it will then show a strong emission band due
to the manganese (maximum 610 nm), together with a weak band charac-
teristic for the lead (maximum 330 nm). The lead absorbs the exciting radia-
tion, part of the excitation energy is emitted by the lead itself, but most of it
is transferred to the manganese. Lead is called the sensitiser, manganese the
activator. Energy transfer from the sensitiser to the activator only occurs
when they are close enough together, although they need not be direct neigh-
bours. This mechanism is responsible for the fact that the ratio between the
intensities of sensitiser and the activator emissions is determined, in the first
instance, by the absolute concentration of the activator, and not by the pro-
portion of the amounts of sensitiser and activator, as might be thought. This
may be shown as follows. Consider a host lattice containing a number of
activator ions. These activator ions may be considered as the centres of
spheres inside of which the sensitiser ions should be situated in order to be
able to transfer their energy to the activator ions. Sensitiser ions outside these
spheres are unable to transfer their energy and can therefore only give their
own emission. According to this picture the ratio between the sensitiser ions
giving their own emission, and those transferring the excitation energy to
the activator, is given by the ratio of the volumes outside and inside the
spheres round the activator ions. This ratio is determined by the concen-
tration of the activator only and not by the concentration of the sensitiser.
This mechanism was found to apply in various sensitised phosphors, albeit
that the picture is sometimes more complicated, e.g. by a transfer from one
sensitiser atom to another.
Little is known about the factors that determine whether energy transport
or energy transfer will lead to luminescence of an activator or not. There
appears to exist a parallel between excitation by cathode-rays and excitation
by irradiation in the host lattice absorption band. Phosphors that have a
high efficiency under cathode-rays are also well excited by very short ultra-
violet, whereas phosphors that can only be excited by irradiation in the acti-
vator absorption band are poor cathode-ray phosphors. Thus, calcium halo-
phosphate, the most important lamp phosphor, is well excited by 254 nm
radiation, hardly at all by 73 nm radiation, and is useless as a cathode-ray
phosphor.
Chapter 4

Luminescent substances
j. L. Ouweltjes

4.1 Chemical composition of phosphors


Phosphorescent materials have been known for a long time. In fact they
were recognised as such long before the phenomenon of :fluorescence was
observed. This is not accidental: fluorescence is only clearly recognised with
the help of special devices such as an ultra-violet lamp, whereas phosphores-
cence may occasionally be observed even after excitation by natural sunlight.
Apart from a vague reference indicating that the Romans may have known
phosphorescent materials, the first case in which this type of material has
been clearly described is that of a Japanese priest who around the year 1000
prepared a pigment by burning oyster shell with sulphur. A painting of an
ox, executed with the pigment on a white background, was invisible in day-
light, but glowed in the dark.
In Europe, the first to make a phosphorescent material was Casciarolo, an
Italian shoemaker and alchemist, by heating barytes with carbon (17th cen-
tury).
The systematic study of phosphors did not start before the end of the
nineteenth century. Among the pioneers the names of Becquerel and Lenard
should be mentioned. The last decades have shown an enormous increase
in the number of phosphors. The application in fluorescent lamps and in
42 FLUORESCENT LAMPS

television tubes aroused the interest of the industrial laboratories and since
then a continuous flow of patents and publications dealing with new phos-
phors and their properties has appeared.
Phosphors are to be found in many varieties. Some occur freely in nature,
e.g. the mineral fluorite (CaF 2 ), from which the term fluorescence has been
derived. Others are willemite (a zinc silicate), scheelite (calcium tungstate)
and certain zinc blendes (zinc sulphides). The luminescent materials used in
lamps are all synthetic, however.
In some of these fluorescence is most pronounced when the substance is
as pure as possible; others do not show any fluorescence unless a small
amount of a foreign element is present. Uranium compounds often fluoresce
in the pure condition, as do various tungstates, such as those of calcium and
magnesium. An example of the other type, which is much more common, is
willemite, which contains a small amount of manganese. As already men-
tioned in the preceding chapter, such an element necessary to induce lumines-
cence is called an activator. The required quantity of the activator varies
within fairly wide limits: for zinc sulphide activated by silver or copper, a
few thousandths of a percent will suffice, whereas the phosphor used in
warm-white lamps (a halophosphate) contains as much as 5% activator. In
general, the efficiency passes through a maximum as the amount of activator
is increased. To indicate the chemical composition of a phosphor it is com-
mon practice to write first the formula of the matrix crystal and then that
of the activator, thus: ZnS : Cu for zinc sulphide activated by copper,
Zn 2 Si04 : Mn for zinc silicate activated by manganese, etc.
Once the leading principle of activation had been properly established,
many new synthetic phosphors were discovered, especially since they proved
to be a vital part of technical devices. Nevertheless, the number of phosphors
used in fluorescent lamps is relatively small. This, of course, is largely due
to the fact that the phosphors have to meet a number of additional require-
ments in order to be suited for their use in fluorescent lamps. Moreover, at
any given time, only these phosphors are used which have the most attractive
properties known. Several of the phosphors in use ten years ago are now
obsolete, and nobody knows whether the phosphors now in use will not be
replaced by others even in the near future.

4.2 Requirements for the practical application of phosphors

4.2.1 The phosphors should have a sufficiently strong absorption for the ultra-
violet rays produced in the particular lamp.
When radiation passes through a liquid or another homogeneous material,
in which it is gradually absorbed, the intensity decreases according to the
formula

in which Id = intensity at penetration depth d, ! 0 = intensity of impinging


radiation and a = a material constant, the absorption coefficient.
LUMINESCENT SUBSTANCES 43
The same formula holds for the absorption of ultra-violet radiation in a
phosphor coating, albeit that the grain structure of the coating causes a
scattering of the rays and therefore an increase of the distance actually
covered by them. Neglecting this scattering for the moment we will calculate
how thick the phosphor coating should be in a fluorescent lamp.
In order to get a good lamp efficiency, only a very small percentage of the
254 nm radiation should be allowed to pass into the glass wall of the tube
unabsorbed. Id/10 is therefore of the order of I%. For IX= 4 600 cm-1,
d is calculated to be 10 f.Lm. For a phosphor having a density of 3, about
3 mg/cm 2 would suffice to absorb 99% of the 254 nm radiation. For a
phosphor having an IX of I 500 cm- 1 about 10 mg/cm 2 would be necessary.
The scattering in the powder and, moreover, the fact that most of the
rays strike the phosphor at a large angle of incidence, results in a stronger
absorption in the coating than would be expected from the formula.
Most phosphors only occur as small crystals, with a particle size of
1-30 flm. It is difficult to measure IX for such materials, and in fact we do
not know the precise values for the various phosphors. The most obvious
way of measuring the absorption coefficient would consist in determining
the transmission for varying coating thicknesses. However, the particle size
of phosphors is in general such that in a thin coating more of the ultra-violet
escapes through openings between the grains than is transmitted by the
grains. Nevertheless is has become clear that there exist important differ-
ences in the values of the absorption coefficients of lamp phosphors. Magne-
sium tungstate has a much stronger absorption than the halophosphates.
For a sufficiently fine magnesium tungstate, one milligram of phosphor per
cm 2 may give the maximum efficiency in a lamp; for a halophosphate the
maximum light output is obtained at much higher coating weights, even if
the particle size is the same. In commercial halophosphate lamps 5-10 mg
of phosphor is applied per cm 2
With halophosphates, moreover, it has been found to be advantageous to
use particles of a larger diameter, even when two coatings are compared
which are both impermeable to ultra-violet. This effect, that has only recently
been recognised, is connected with the backward reflection of the ultra-violet
by the phosphor.
When radiation passes the surface that separates one medium from an-
other, part of it is reflected. For a powdered material the number of reflecting
surfaces is very large. In a white powder no absorption occurs, so that in an
infinitely thick layer all of the radiation must, in consequence, be ultimately
reflected backward and escape again from the surface. If, on the other hand,
the radiation is absorbed completely on its way through one grain, the reflec-
tion will be restricted to those surfaces struck directly by the radiation. Such
a material may be said to be black for the radiation concerned. The ideal
lamp phosphor is white for visible light, but black for 254 nm rays. If ab-
sorption in a single grain is not complete, the material may be said to be
gray. Whether a material is black or not depends not only on the absorption
coefficient IX, but also on the particle size d. A phosphor may be called black
for 254 nm radiation if e-ad is of the order of a few percent.
For halophosphates IX is of the order of 1 500 cm- 1 From the absorp-
tion formula we find that, for Id = 5% of 10 , d = 20 f.Lm. If we consider
44 FLUORESCENT LAMPS

the case of cubic particles, and remember that the radiation reflected from
the bottom surface has to pass through the crystal twice, then it may be
expected that, for a particle size of 10 f.Lm and less, reflections from these
bottom planes will begin to contribute to the amount of 254 nm radiation
reflected by the phosphor. This proves indeed to be the case. With decreasing
particle size the reflection rises slowly at first, but more rapidly as the particle
size is further reduced. The following table of measurements on various
fractions from one halophosphate phosphor may illustrate this effect.

Particle size, average 115 f.Lm 60 f.Lm 46 f.Lm I32 f.Lm 16 f.Lm
Reflection I 8% 10% 15% 1 2o% 38%

From experiments with phosphors of varying reflectivity it is estimated


that about one third of the reflected 254 nm radiation is lost in the lamp.
A lamp made with the 60 f.Lm fraction may be expected, therefore, to give
about 9 % more light than a lamp coated with the phosphor of 16 f.Lm par-
ticle size.
Preparation of the halophosphate phosphors in such a way that they are
practically free from particles smaller than 5 f.Lm has indeed resulted in a
remarkable improvement in the lamp efficiency.

4.2.2 The conversion of the absorbed ultra-violet radiation into visible light
should be as efficient as possible
In this connection the question may be raised as to how the efficiency
should be indicated, and also what may be expected to be the maximum
efficiency. Both questions may be answered by a closer consideration of the
mechanism of luminescence discussed in Chapter 3.
Irrespective of the details, which are different for various types of phos-
phors, for each quantum of ultra-violet radiation absorbed by the phosphor,
a quantum of visible light may be emitted instead. This is the ideal case. In
practice part of the exciting radiation may be absorbed by impurities or
crystal defects, and will be lost for the purpose of excitation. Moreover, not
all of the excited activator atoms must necessarily return to the ground state
by means of luminescence. Radiationless transitions are always possible and
become increasingly important as the temperature is raised.
The best efficiency to be expected corresponds, therefore, to one quantum
of visible light for each absorbed quantum of ultra-violet radiation. The
quality of a phosphor may thus be expressed in terms of the quantum effi-
ciency, which gives us the ratio of the number of emitted quanta to the
number of absorbed ultra-violet quanta.
Most phosphors used in fluorescent lamps have a quantum efficiency of
about 80%.
In Chapter 1, we have also seen that each quantum of radiation corre-
sponds to an energy: E = hv. The energies, therefore, are inversely propor-
LUMINESCENT SUBSTANCES 45
tional to the wavelength. If a phosphor absorbs 254 nm radiation and emits
visible light with a wavelength of 508 nm, than each quantum of the visible
light will have half the energy of the absorbed quantum. As a consequence,
even in the case of maximum quantum efficiency, half of the energy will be
lost. This energy is dissipated in the form of heat.
The contents of Chapter 6 will make it clear that the quantum efficiency
of a phosphor is not equivalent to the quantity of light produced by it. The
human eye is more sensitive to yellowish green light than to red or blue
light. Red and blue emitting phosphors will therefore always be less efficient
in light production, even if the quantum efficiencies are the same.

4.3. Physical properties of phosphors


4.3.1 Zinc silicate and zinc beryllium silicates
One of the earliest known phosphors used in gas discharge devices was
zinc silicate, which, when activated with 02-2% manganese shows a bright
green fluorescence. The green neon signs and the green oscillograph tubes
are still made with this zinc silicate, often called willemite in analogy to the
mineral. For fluorescent lamps it is still used mainly as a means of adjusting
the colour point. Willemite has a narrow emission band with a peak at
525 nm (see Fig. 4.1).

450

Fig. 4.1. Spectral energy distribution of zinc silicate and zinc beryllium silicate.

Substitution of part of the zinc by beryllium leads to a group of phosphors


that was universally used for fluorescent lamps up to 1948-1950. The zinc
beryllium silicates have a second band at 610 nm next to the 525 nm band
observed for willemite. If the manganese content is increased this last men-
tioned band increases at the expense of the green one. In this way it is pos-
sible to prepare a range of phosphors of varying colours having a good
efficiency and a spectral distribution very well suited for use in white fluores-
cent lamps. The zinc beryllium silicates have been abandoned since 1950
46 FLUORESCENT LAMPS

because of occasional health hazards encountered by those working in the


production of phosphors and lamps, and the safety problems arising in the
disposal of the lamps.

4.3.2 Magnesium tungstate and calcium tungstate


Magnesium tungstate has a broad emission band extending almost over
the whole of the visible spectrum, with a flat topped maximum at 480 nm.
The colour of the emitted light is bluish white. Together with yellow phos-
phors magnesium tungstate is particularly well suited for making white lamps
and until the introduction of the halophosphates it was universally used for
this purpose. Most lampmakers now use a blue halophosphate. Calcium
tungstate has the same type of emission spectrum, the whole being shifted
to shorter wavelengths. The maximum emission lies at 410 nm; by adding a
small quantity of lead it may be shifted to 430 nm. The variety without lead
has often been used in lamps for blue-printing. The lead-activated phosphor
may be used to make deep blue lamps, or as an addition to a phosphor
mixture for lamps in which we wish to increase the amount of deep blue.
The tungstates are different from the majority of phosphors in so far as no
activator is present, the absorption and emission taking place in the tungstate
groups.

4.3.3 Halophosphates
Since the use of zinc beryllium silicate was abandoned, the halophosphates
have become the most important phosphors for low pressure mercury lamps.
As compared with the old lamps containing zinc beryllium silicate the
luminous output is higher and the maintenance during lamp-life is also
better. The higher light efficiency is due to the fact that the amount of light
radiated for the average quantum is greater than for the zinc beryllium sili-
cate-containing phosphor mixtures having the same colour. The chemical
composition is very similar to that of the natural mineral apatite: 3Ca(P04 ) 2 -
CaF 2 Most of the phosphors used in lamps also contain some chlorine,
which replaces part of the fluorine. Part of the calcium is substituted by
antimony and manganese, both of which serve as activators. The antimony
produces a blue emission, very similar to that of magnesium tungstate, the
manganese a yellow-orange band. The position of the manganese band de-
pends on the fluorine-chlorine ratio and on the manganese content. Both
the increase of the chlorine content and the manganese content result in a
slight shift of the emission towards longer wavelengths. In this way materials
may be prepared having a peak wavelength varying from about 570-595 nm.
The halophosphates belong to the class of sensitised phosphors, already
mentioned in Chapter 3. The exciting radiation is absorbed by the antimony
ions. Some of these transfer their excitation energy to manganese ions in their
vicinity, which in turn become excited and return to the ground state under
emission of the manganese band. As was explained in Chapter 3 this
mechanism is responsible for the fact that the intensity ratio between the
LUMINESCENT SUBSTANCES 47
antimony and the manganese bands is determined, in the first instance, by
the absolute manganese content, and not by the proportion of the amounts
of antimony and manganese. By varying the manganese content the colour
changes from blue (no manganese) over blue-white and white to yellowish
white. This makes it possible to make the whole range of existing white
lamps (daylight, cool white and warm white) with halophosphates. In general
the cool white and warm white phosphors are made such as to yield the
correct colour points of the lamps; for making daylight lamps either of these
phosphors maybe mixed with a blue halophosphate or magnesium tungstate.
Fig. 4.2 shows the emission curves of three halophosphates.

400 500 600


- - },(nm)

Fig. 4.2. Spectral energy distribution of three halophosphates. (a) blue halophosphate;
(b) halophosphate 4200 K; (c) halophosphate 2900 K.

4.3.4 Lead-activated silicates


Among the alkaline-earth &ilicates there are several which may be activated
by lead and in this way give rise to efficient ultra-violet phosphors. Three
applications are especially of interest: for erythemal lamps (A about 300 nm),
for 'black light' lamps (A about 350 nm) and for blueprinting lamps (A about
380 nm). For all three applications suitable phosphors have been developed.
For the erythemal region barium zinc silicate may be used. The peak of
its emission curve (298 nm) is very close to the peak in the erythemal activity
curve (297 nm). Barium disilicate has an emission peak at 350 nm and is at
the present the most widely used phosphor for 'black-light' lamps. For blue-
printing lamps, ternary silicates (barium-strontium-magnesium silicate or
barium-strontium-zinc silicate) have until recently been the most efficient
phosphors known. The emission curves of the various phosphors are given
in Fig. 4.3.
48 FLUORESCENT LAMPS

BZS 805 SMS

t
250 450

Fig. 4.3. Spectral energy distribution of lead-activated silicates.

Lead may also act as a sensitiser. The calcium silicate activated by lead and
manganese that was discussed as an example of sensitised luminescence in
Chapter 3, is the most important example. It has long been in use, and is
still being used by some lampmakers in the phosphor mixture for the de
Luxe lamps (see Chapter 6).

4.3.5 Magnesium arsenate and magnesium germanate activated by manganese


These phosphors are very similar in their properties. Both have a sharp
peak in their emission curves at 656 nm (see Fig. 4.4) and in consequence
display a remarkable bright red fluorescence, under both 254 nm and 365 nm.
This together with an exceptionally good temperature-dependence, makes
them also attractive for use in high pressure mercury lamps, which are rich
in long wavelength ultra-violet. Whereas the other phosphors are white in
appearance in daylight, magnesium arsenate and germanate are yellowish,
due to a weak absorption of the deep blue radiation. Since this adsorbed blue
radiation is converted into red light, this opens an attractive possibility for

~ 100
j
~

i 50

400 500 600 800


-;\(nm)

Fig. 4.4. Spectral energy distribution of magnesium arsenate: manganese (a) and of
strontium magnesium orthophosphate: tin (b).
LUMINESCENT SUBSTANCES 49
reducing the intensity of the blue mercury line. This will be discussed in
Chapter 6. Chemically, it is interesting that in these phosphors the manganese
is tetravalent. In this state of valency the energy levels involved in the optical
transitions are more or less shielded from external influences; as a conse-
quence the emission spectrum consists of a number of very narrow sub-bands,
which are very similar for all materials in which Mn4+ is an activator.

4.3.6 Alkaline earth phosphates activated by tin


Tin acts as an activator in various phosphates, but so far only the earth
alkaline orthophosphates have gained practical importance. Calcium ortho-
phosphate: tin was the first to be introduced as a lamp phosphor; at the
moment the strontium magnesium orthophosphate is almost universally used
by the lamp makers. It has a broad emission band with a peak at 630 nm,
and a second low intensity band in the near ultra-violet (see Fig. 4.4). The
quantum efficiency (,..., 85 %) is considerably higher than that of the cal-
cium silicate phosphor aforementioned (,..., 65 %), but the luminous effi-
ciency is hardly any better, due to the deep red and infra-red radiation.
Still the large amount of deep red radiation makes it an attractive phosphor
for lamps with good colour rendering properties.

4.3.7 Europium activated phosphors


Phosphors activated by the so-called rare earth elements (cerium, sama-
rium, europium) have been known for a long time, but until recently they
were considered as being very expensive without having specific advantages.
This all changed when europium activated compounds (yttrium oxide,
yttrium vanadate and others) proved to be particularly well suited for the
red component in colour television tubes. They all have an emission spec-
trum consisting of extremely narrow bands, the one with the highest intensity
lying at about 615 nm. In these phosphors europium is in the trivalent state.
Divalent europium can also act as an activator. Incorporated in alkaline
earth ortho- or pyrophosphates, or in alkaline earth silicates, it produces
emission bands in the near ultra-violet or in the blue part of the spectrum.
The strontium orthophosphate and strontium pyrophosphate phosphors
have already proved their usefulness in lamps for photo-copying, whereas
others may find application in the lamps with very good colour rendering
properties.

4.4 Preparation of phosphors


In general, phosphors are prepared by heating together a mixture of suitable
ingredients at a temperature high enough to cause a chemical reaction, lead-
ing to the formation of the desired compound. This is usually done by
putting the mixture in a silica crucible or tray and firing in an electric furnace
50 FLUORESCENT LAMPS

at I 000-1 200 C. Nearly all phosphors used in lamps are salts of oxygen
containing acids, such as the silicates and phosphates. In most cases it is
impossible to prepare these compounds by the classical method of precipi-
tation. Even if it were possible, we prefer to prepare the matrix by firing a
mixture of basic and acidic components. In the case of the silicates, silica
may be fired together with the oxide or the carbonate of the metal in question.
For the preparation of zinc silicate, zinc oxide is generally employed; the
silicates of the alkaline earth metals (Ca,Sr,Ba) are prepared by heating silica
together with the corresponding carbonates.
The phosphates may be prepared in a variety of ways. The commercial
halophosphates are generally made by heating together calcium hydrophos-
phate (CaHP0 4 ), calcium carbonate, calcium fluoride and calcium chloride
as the constituents forming the matrix, whilst the activators antimony and
manganese are added in the form of the oxide and the carbonate respectively.
It must be emphasised that the firing temperature in the synthesis of phos-
phors remains well below the melting point of the material to be formed.
The chemical reactions obviously take place in the solid state. In the last
few decades there has been a lively interest in this type of reaction, which is
also of importance in the chemistry of ceramics etc. It has become clear that
the ions or atoms in a solid change their sites in a crystal lattice at tempera-
ture far below the melting points. In this way chemical reactions may occur,
which have the character of a diffusion process. In the synthesis of zinc
silicate from zinc oxide and silica, the zinc diffuses into the silica, forming
a thin layer of zinc silicate. After this layer is formed, the reaction proceeds
by the diffusion of zinc ions through the zinc silicate, causing this layer to
increase in thickness until all of the zinc oxide is used.
In the case of the synthesis of halophosphates, the mechanism is more
complicated because of the many ingredients used. The fundamental pattern
is nevertheless very much the same: the metal ions diffuse into the acidic
constituent, in this case the calcium hydrophosphate.
We know of only two cases in which the acid radical seems to be the diffu-
sing agent: the formation of calcium tungstate and magnesium tungstate
from tungstic acid and the corresponding metal oxides and the synthesis of
magnesium arsenate from magnesium oxide and arsenic acid.
It may be asked what happens to the activator. In Chapter 3 it has been
pointed out that there is nothing special about the site of the activator. the
ions or atoms occupy normal lattice positions, and from this point of view
no special behaviour should be expected in the synthesis of phosphors. In
most cases the activator ions will diffuse into the acid constituent just like
the ions of the bulk metal. There are, however, a few details that should be
kept in mind. The activator remains a foreign constituent in the matrix lat-
tice. Thus the incorporation will in general be more difficult than the incor-
poration of the matrix ion. As a consequence, in the presence of an excess
of the matrix metal oxide, there is a chance that the activator will not be
incorporated at all, so that the fired material is non-luminescent. This is one
of the reasons why the composition is chosen so that there is a slight excess
of the acidic oxides.
In some cases the valency of the activators offers a problem. In calcium
silicate activated by lead and manganese, the incorporation of manganese
LUMINESCENT SUBSTANCES 51
is easiest in a reducing atmosphere, but then the lead oxide will be reduced
to the metal. In an oxidising atmosphere the lead offers no difficulty, but now
the manganese is easily oxidised to undesirable peroxides. A similar problem
is encountered in the synthesis of tin-activated alkaline earth phosphates. In
air, all of the tin remains tetravalent, whilst in a strongly reducing atmosphere
reduction to the metal occurs. Since only the divalent tin serves as an acti-
vator, the atmosphere during the firing process is very critical.
After the firing process the phosphor forms a sintered cake, that can be
broken up in a crusher and reduced to particles having a diameter of 1-20 [Lm
by milling in a ball mill. One of the problems of phosphor synthesis is to
avoid heavy sintering in the firing proces. This would make it necessary to
apply heavy mechanical forces in milling and this often will result in a
serious drop in efficiency of the phosphor. Most of the phosphors are sensi-
tive to crushing and grinding, some of them being so sensitive that special
techniques have to be developed in order not to lose too much in efficiency.
Among them is cadmium borate activated by manganese. It has long been
in use in fluorescent lamps, but has been replaced by other phosphors, one
of the reasons being this sensitivity to crushing.
The purpose of industrial phosphor production is to make materials having
maximum efficiency. An important point is the chemical purity of the ingre-
dients used for the synthesis, especially as the presence of heavy metals such
as iron, nickel and the like often causes a drop of the efficiency; these ele-
ments absorb the exciting radiation and transform it into heat. As a general
rule the allowable amount of harmful elements is about 1 % of the activator
content. For most of the lamp phosphors, which contain 01-1% activator,
this means that a chemical purity is required which corresponds to that used
in analytical chemistry.
For the sulphide phosphors the activator concentration is much lower, and
in consequence the purity requirements are correspondingly higher. The zinc
sulphide used for the production of TV phosphors should have not more
than w- 5 % of undesired elements. Apart from the purity of the ingredients,
the particle size is very important, since the particle size determines to a large
extent the speed of the chemical reactions. The finer the ingredients, the
quicker the reactions proceed.
Despite the vast amount of experience already gained it is still impossible
to predict whether a specific combination of host crystal and activator will
provide a good phosphor or not. As a consequence, the search for new phos-
phors is largely empirical.
Chapter 5

Lamp design and lamp manufacture


L. E. Vrenken

5.1. Essential parts


As pointed out in the previous chapters the fluorescent lamp is an electric
discharge source of the low pressure mercury vapour type in which light is
produced predominantly by the fluorescence of a translucent coating of
fluorescent powder excited by ultra-violet radiation from the mercury vapour
discharge. The best conditions under which the desired ultra-violet radiation
occurs are low current density and low vapour pressure (see Chapter 2).
The lamp therefore is usually in the form of a relatively long tubular bulb
with an electrode sealed into each end and contains mercury vapour at low
pressure with a few torr of a rare gas. The inner wall of the bulb is coated
with a phosphor or a mixture of phosphors. The lamp must be connected
to a ballasted electrical circuit by means of a cap fixed onto each end of
the lamp. The essential parts of the lamp are therefore bulb, electrodes,
mercury and rare gas, phosphors and caps.
The phosphors themselves are dealt with in Chapter 4. The coating process
and the optical properties of the phosphor coating will be treated in this
chapter.
LAMP DESIGN AND LAMP MANUFACTURE 53
5.2. Bulbs
Fluorescent lamps are most commonly made with tubular bulbs varying
in diameter from 16-54 mm and in overall lengths from a nominal 15-240
em. In the U.S.A. the bulb is usually designated by a letter indicating the
shape followed by a number indicating the maximum diameter in eights of
an inch. Hence 'T-17' indicates a tube of 17/8 in diameter.
Tubular lamps in the form of a circle (circline), in U-shape, or using inter-
mittently grooved tubes are also manufactured. As is pointed out in Sections
2.1 and 2.2 there is an optimum bulb-diameter for every value of the current.
For standardisation reasons the following diameters for the main straight
types of lamps have been internationally agreed upon.

Nominal value Normal current load


16 mm 5/8 in (T5) !SOmA
25 mm 1 in (T8) 300mA
38 mm It in (T12) 400--1500 rnA
54 mm 2t in (Tl7) 1500 rnA

These diameters are not optimal for all the current values they are used
for. For some currents they are too wide, for other currents they are too
narrow. Preferred lengths are 24, 36, 48, 60, 72 and 96 in, although other
lengths also occur. For straight tubes sodalime glass is the most practical
and cheapest material, but for circlines and other more complicated shapes
a leadglass, with a longer melting zone has many advantages. As it is much
more expensive than sodalime glass the latter is applied more and more for
these lamps too. For some applications fluorescent lamps are used as ultra-
violet radiators. In that case the bulb may be made of a special glass, which
transmits this radiation.

5.3. Electrodes
The function of the electrode is to facilitate the transfer of the current
from the metal conductor to the gas, and vice versa. Here the difficulty
mainly concerns the transfer from metal to gas, since electrons will easily
enter a metal (releasing energy) but cannot so easily leave it (see Section 1.2).
When operating on alternating current, each electrode functions alterna-
tely as a cathode and as an anode, and both ends of the lamp are therefore
identical. Although there are two basic types of fluorescent lamps - hot
cathode and cold cathode lamps - most fluorescent lamps are of the hot
cathode type. The design of the electrode is governed mainly by its function
as cathode. The electrodes are coiled-coil or triple-coil tungsten filaments
coated with one or more of the alkaline earth oxides. Tungsten is used be-
cause other metals evaporate too rapidly, resulting in blackening at the ends
of the lamp.
The electron-emissive coating, when hot, provides an abundance of free
electrons. By suitable circuit arrangements the cathodes can be heated to a
temperature high enough to warrant good electron emission before the arc
54 FLUORESCENT LAMPS

is struck (preheat, TLM or rapid start) or by the ion bombardement after


the arc is struck (instant start, TLS or slimline).
The dimensions of the electrode are determined by the lamp current, the
filling gas and by the method of starting (see Section 1.8).
An electrode of a lamp which uses preheating has to be so dimensioned
that the preheating current is sufficient to heat the cathode above the so-
called emission temperature.
Especially in capacitive circuits, where as a rule the short circuit current
approximately equals the lamp current, the diameter of the tungsten wire
should not be too large.
On the other hand for the sake of lamp efficiency and wattage loss in the
electrodes during life this diameter must not be chosen too small. The reason
is that in the course of lamp life the hot spot of the discharge shifts very
slowly from the connected end of the electrode to the opposite end, caused
by the consumption of the emitter material. This means that during life the
electrode losses are increased with the product of lamp current and the
voltage across that part of the electrode, through which the lamp current
flows.
The minimum preheating current for the cathode can be read from the
so-called emission curve, indicating the relation between ignition voltage of
the lamp at 25 oc ambient temperature and the heating current through the
electrodes. It appears that the ignition voltage falls rapidly and considerably
according as the temperature of the electrodes is increased. As soon as the
electrodes supply a satisfactory thermal emission, the ignition voltage hardly
diminishes any longer, so that from that point on the curve runs nearly
horizontally.
Fig. 5.1 shows this relation for a typical 40 W lamp. See also Fig. 1.12.
The current at point e of the curve represents the minimum required pre-
heating current I . For this particular electrode this current is 315 rnA.
For all circuits in which the two electrodes of the lamp are heated in
series with the ballast and where therefore the latter determines the pre-
heating current, the above-mentioned emission current only is decisive for

225
s:
'-
c:
-;;..':!.>200
\
i 175 \
\.e
150

125
225 250 275 300 325 350 375 ~()()
- Preheating current (mA)

Fig. 5.1. V19 as a function of the preheating current for a 40 W fluorescent lamp with a
certain electrode. At point e the 'emission temperature' is reached. For a good ignition
condition the supplied preheating current must be well above the value I .
LAMP DESIGN AND LAMP MANUFACTURE 55
the cathode design. The voltage across each electrode is not important.
In many starterless circuits, however, both electrodes are heated by means
of transformer windings, providing a voltage across each electrode. With
this voltage sufficient preheating current must flow. Two groups of lamps
can be distinguished here, those with low resistance cathodes and those with
high resistance cathodes.

Lamps with low resistance cathodes


At the nominal mains voltage the ballasts supply a heating voltage to
either electrode of about 4 V. According to I.E. C. publication 82 * the star-
terless 40 W ballast for instance should provide a heating voltage of at least
305 V at 90% of the rated supply voltage and with a 10 n substitute resis-
tance across the preheat transformer windings. This means for the electro-
des, that at a voltage of 305 V a preheating current must occur which at
least equals the current Ie (see Fig. 5.1). For the protection of the trans-
former windings of the ballast it has, moreover, been laid down in I.E.C.
Publication 81 that the electrode resistance at a 36 V test voltage should
amount to at least 7 n. From the voltage-current characteristic of the elec-
trode it can be determined, whether these two requirements are met. In

/ "'
e /
v
--
v
.............

0
225 250 275 300 325 350 375 1,()()
- - Preheating current (mA)

Fig. 5.2. Current-voltage characteristic of a certain 40 W fluorescent lamp electrode. The


minimum required preheating current and preheating voltage for good ignition of the
fluorescent lamp are for this electrode 315 rnA and 30 V respectively.

Fig. 5.2 this voltage-current characteristic is given for a certain 40 W lamp


electrode. It can be seen, that at 305 V the minimum required preheating
current of 315 rnA is just reached. Furthermore it can be concluded from
the figure that at 36 V the electrode resistance is much higher than the
required minimum of 7 n. In I.E.C. publication 81 minimum values for
other lamp wattages are given.

* I.E. C.: International Electrotechnical Commission.


56 FLUORESCENT LAMPS

Lamps with high resistance cathodes


Here a heating voltage is required of at least 8 V at 90% of the mains
voltage with for instance for the 80 W lamp a 12 n substitute resistance
across each of the transformer windings, whilst the coil resistance at a
heating voltage of 8 V should amount to at least 9 n. In I.E.C. publication
81 minimum resistance values for other lamp wattages are given.

The above mentioned requirements for the electrodes relate to the ignition
condition of fluorescent lamps. But in the situation of norma;l burning of the
lamp in starterless circuits the maximum current in any of the four leads to
both electrodes too is of importance. The maximum current in one of the
four leads is decisive for the maximum temperature that will occur at the
corresponding coil extremity. Consequently, the blackening and the life will
also be determined by this maximum lead current.
In Fig. 5.3 the current distribution in an electrode is shown schematically
by assuming, for the sake of simplicity, that the lamp current 1, starts from
a certain point of the electrode.
-+
The two lead currents are / 1 and / 2 , which generally speaking are not in
-+
phase mutually and with h The resistance of the electrode at the prevailing
temperature conditions is R 1 + R 2 = R.

-I
It I
Fig. 5.3. Current distribution at one lamp end; /7 and J;
are the lead currents; 7, is the lamp current. R 1 and R 2
are the resistances of the electrode parts.

As a result of the currents passing through the electrode when the lamp
-+
burns, a potential difference Ve 1 occurs across the electrode. From the two
vector equations

and -+
/1R1- l2R2 = Vel
it follows that for the lead current:

~ 1 -+
+ (1- p) I,
->
/2 = - - Vel
R
in which p = R 2 /R (0 ~P ~ 1). See also Fig. 5.4.
LAMP DESIGN AND LAMP MANUFACTURE 57

pi(
I;
Fig. 5.4. Vector diagram of the lamp current
~ ~
J;, the potential difference V.,, across the
electrode, and the lead currents / 1 and / 2

-> -;.
When 11 and Ve 1 are given, the greatest value of the maximum lead cur-
__,.
rent occurs when p = 1 and Ve 1 and 11 are in phase, or when p = 0 and
~

Ve 1 and / 1 are in anti-phase.


Vel
In the former case Imax = / 1 = 11 + - = I 1 + /heat and
R

in the latter case

The smallest value of the maximum lead current (lm 1n) occurs when
-+ -> -+
/ 1 = / 2 (thus equal to 1- / 1).
1 ~ ~ 1 ..... .....
This is the case when- Ve 1 + p 11 = - - Ve 1 + (1- p) I,
R R
so when V:,, = R (!- p) i;. Then Imin will be I 1 = I2 = tf1
The lamp current is then equally divided over the two leads to one lamp
end. From the foregoing it follows that the maximum lead current always
lies between the values 1- 11 and 11 + /heat Whether this value is admissible
for the maximum lead current of a fluorescent lamp in a starterless circuit
can only be determined eventually by means of lamp life tests.

5.3.1 Geometry of the electrode:


The tungsten electrode may be of the coiled coil or of the triple coil type
(see Fig. 5.5).
Although a triple coil is more expensive than a coiled-coil electrode the
former has so many advantages above the latter, that it has found a wide-
spread use. The required amount of emitter can be applied more easily, and
the triple coil is much more resistant against cold starting, resulting in less
end blackening and longer life. The geometry of the triple coil (diameters
of tungsten wires and mandrel pitches) determines lamp life, because it de-
termines to a high degree the amount of emitter which can be applied without
58 FLUORESCENT LAMPS

Fig. 5.5. Schematic drawing of a triple coil electrode. A thin tungsten wire A is loosely
wound around the main tungsten wire B. The combination of A and B is coiled twice.

risk of loosening of the emitter from the electrodes. The required amount of
emitter per 1000 hrs life is a.o. strongly dependent on nature and pressure
of the filling gas, this amount decreasing with heavier rare gas and with
higher pressure. As already stated, the electrode which functions as cathode
in one phase is virtually the anode in the next, and for this reason the design
of the electrode is sometimes more complicated than simply a coated coiled-
coil or triple coil.
Examples of electrode design are given in Fig. 5.6.

Fig. 5.6. Projection of the electrode on a plane perpendicular to the lamp axis. (a) Simpler
coiled coil or triple coil as electrode; (b) As a but with twowiresorplatesparalleltothe
coil; (c) As a but with a 'floating' ring around the coil.

In Fig. 5.6b two wires or plates run parallel to the coil. They serve to
take part of the current during the phase in which the electrode is acting as
anode. In Fig. 5.6c a ring of about 6 mm height is mounted around the
coil, commonly on a separate support, so that it is not connected with the
LAMP DESIGN AND LAMP MANUFACTURE 59

Fig. 5.7. Coated electrodes.

electrode. This 'floating' ring prevents material being sputtered away or


evaporated from the electrode from reaching the coated end of the lamp and
therefore is a protector against end blackening. Moreover the ring reduces
the anode fall and thus improves the luminous efficiency significantly.
Another advantage of the floating ring is that it reduces by a considerable
amount end flicker in fluorescent lamps.
We have already pointed out that the electrode coils are coated with
oxides of barium, strontium and calcium in order to reduce the work func-
tion. As a rule, this emitter is applied in the form of Ba-Sr-Ca-carbonate
which is decomposed during the pumping process to give Ba-Sr-Ca-oxide
and C0 2 , the latter being pumped or flushed away during this process (see
Section 5.9). The carbonate can be applied by dipping, spraying or by cata-
phoresis. Fig. 5. 7 shows a photograph of mounts with coated electrodes
before decomposition.

5.4 Mercury and rare gas


The influence of the mercury pressure is dealt with in Section 2.5.1. From
there it follows that the exact amount of mercury is not important. A quan-
tity of 10 mg may be considered as a safe minimum. Overdosing has to be
avoided, for this may lead to mercury condensation spots on the phosphor
coating.
For some types of lamp the exact amount of mercury is important, namely
for typical amalgam lamps, where the amalgam is formed in the lamp after
lamp making. An alloying metal such as indium is placed in the bulb before
sealing in; a practical solution is, for example, a small indium disk pressed
onto the tube wall before the phosphor coating is applied. As the indium/
mercury ratio and the temperature determine the mercury vapour pressure,
an exact dose of mercury is required in order that the lamp has the maximum
luminous flux at a given ambient temperature. Besides this mercury vapour
regulating mechanism a quick supply of mercury is necessary in amalgam
lamps to increase rapidly the very low mercury vapour pressure after ignition
60 FLUORESCENT LAMPS

of the lamp to the required 5 X I0- torr. A practical solution here is to


3

apply additional indium on the floating ring. This ring is heated up rapidly
after ignition of the lamp, and the amalgam on the ring delivers enough
mercury for the lamp to reach full light output quickly. This effect is shown
in Fig. 5.8 for typical 40 W amalgam lamps with and without indium on
the ring in comparison with normal40 W lamps. Note also that in this closed
four lamp fixture normal lamps deliver only approx. 75% of the light that
amalgam lamps give.

--- ------
--. __

/
/
/
i
Ia
i
o~------------~----~------~----~~----~----
0 10 20 30 . 1.0 . . . 50 60 . I
- Ttme after tgntttOn of lamps (mmutes,

Fig. 5.8. Relative luminous flux from a closed four lamp fixture as a function of the time
lapse after ignition of the lamps. (a) 40 W amalgam lamps without indium on the ring;
(b) 40 W amalgam lamps with indium on the ring; (c) normal 40 W lamps.

Needless to say, the indium on the ring does not interfere with the above
mentioned mercury vapour pressure control, because the temperature of the
ring when the lamp burns is much higher. In Section 2.5.5 the influence of
the pressure and kind of rare gas is dealt with. For example: in 40 W fluores-
cent lamps a filling pressure of 25 torr argon or argon with 25% neon is
very common.

5.5 Caps
To connect a lamp to the electrical circuit a number of different lamp caps
have been designed. The design depends amongst other things on the require-
ments of the circuit. Lamps for switch start, semi-resonance or rapid start
circuits require two connections per electrode, thus a bi-pin type cap is used
at each end, or a recessed double contact cap. For safety reasons the latter
is especially used for some long lamps in circuits with high no load voltage.
For instant start lamps one single connection for each electrode is sufficient.
Here a single pin cap is used for slimline lamps, a single cap of special design
LAMP DESIGN AND LAMP MANUFACTURE 61

Fig. 5.9. Some typical caps for fluorescent lamps:


a. bi-pin cap for T12 lamps;
b. four pin cap for circular lamps;
c. recessed single contact cap for TLS lamps;
d. single pin contact cap with narrow tolerances for TLX lamps;
e. single pin contact cap for slimline lamps;
f. recessed double contact cap for T12 lamps.

for TLX (safety) lamps and a recessed single contact cap for TLS and TLR
lamps. A circular lamp is fitted with one fourpin cap (see Fig. 5.9).
The dimensions of the cap on the completed lamp shall be in accordance
with those given in I.E.C. publication 61.

5.6 Lamp making


The bulbs for fluorescent lamps are coated with a phosphor by flushing
them with a suspension of a phosphor in a binder (see Section 5.7). After
drying of the coating the binder is removed by heating the tubes in a lehr
at 550-600 oc. Then the mounts on which the electrodes are fixed are sealed
onto the bulb, one on each side. At least one of these mounts is provided
with an exhaust tube. While the lamps are either still hot from the lehr or
heated up again, the unwanted gases are removed by an exhaust machine
and replaced by Hg and a rare gas. Moreover the electrodes are degassed
and activated (see Section 5.9). Caps are then fitted at each end of the tube
62 FLUORESCENT LAMPS

after which the lead wires are threaded through the hollow pins in the caps;
a special cement is used for fixing these caps in position and the leads, after
being trimmed to the correct length, are connected to the pins by soldering,
welding or clamping. Due to the fact that the gas in the lamp is not abso-
lutely pure it is commonly necessary to operate the lamp for a short time to
give the fluorescent powder an opportunity to remove the impurities from
the gas by a gettering action. In some cases the manufacturing is more com-
plicated, for instance in the case of the lamp requiring an internal ignition
strip, or where two layers of phosphors are used.
The latter occurs in some of the 'de Luxe' fluorescent lamps (Chapter 6)
or in reflectorised lamps where before being coated with fluorescent powder,
a first layer of reflecting powder is applied, leaving a window at the place
where the lamp has to have a higher luminance (TLF and aperture lamps).
Rapid start lamps have to be provided with a dry film coating of silicone
on the outside of the bulb to prevent poor starting in a humid atmosphere.
We will discuss these processes more in detail in the next sections.

5.7 Coating the tube wall with a phosphor


As mentioned in Section 5.6 the phosphors are generally applied to the
tube wall by flushing the tubes with a suspension of the phosphor in an
organic binder. Commonly used binders are dilute solutions of a high vis-
cosity nitrocellulose or ethylcellulose in butylacetate or xylene. Solutions of
high polymeric substances in water may also be used. The goal of suspension
making is to get an even dispersion of the primary particles throughout the
liquid. Preparation of suspensions by milling in a ball mill is very common,
but easily causes a decrease of the quantum efficiency or an increased ultra-
violet reflection caused by the small particles, both resulting in a lower
efficiency. This is especially important with the halophosphates (see Section
4.2). Therefore the phosphor production aims at a phosphor with the right
grain size from the start, so that it can be dispersed by stirring with a high
speed stirrer without any actual milling.
The phosphor suspension may be introduced by spraying it in from the
top ('down flush') or by forcing it upward into the tubes from a container
by applying air pressure ('up flush'). When the pressure is removed, the sus-
pension flows back into the container, leaving a thin coating of phosphor
on the tube wall. A current of air is blown in from the top of the bulb and
by selecting the right amount of air and the temperature during drying it is
possible to get reasonably even coatings. Water binder suspension requires
drying with hot air from approx. 100 oc, resulting in a relatively short
drying time. In order to obtain an even and smooth coating, which is free
from defects, the drying should be carried out very carefully. The down
flush method especially can lead to high speed fully mechanised coating and
drying units.
LAMP DESIGN AND LAMP MANUFACTURE 63
5.7.1 Requirements for the coating
The objective in applying the coating is to obtain a uniform coating of
good appearance and the proper thickness over the entire inner surface of
the bulb. The light output of the lamp, amongst other things, depends on
the thickness of the phosphor coating. As can be seen in Fig. 5.13 the phos-
phor coating shows holes between the grains. With a thin layer ultra-violet
radiation will fall through the openings in the layer onto the glass and will
be lost for excitation purposes. On the other hand, with a very thick coating
the light which is mainly generated at the inside surface of the coating will
be absorbed to a certain extent when passing the fluorescent coating. The
optimal coating thickness represents a compromise between these two ef-
fects (see Fig. 5.10).

/ I ---l_ ........__
'?3200
<::::.

~
1/
I
Ill
0
:J i
-~ 3750
.....,
I
v
i
I
I

!
3100
40
- 60
50 70 80 90
Powder weighYcm tube length (mglcmJ

Fig. 5.1 0. Luminous flux of a typical 40 W fluorescent lamp as a function of the applied
powder weight per em tube length.

The factors which affect coating thickness are, amongst others, the vis-
cosity of the coating suspension and control of the drying operation. In
order to ensure the highest possible lamp efficiency the coating thickness
should therefore be checked constantly. This can be effected in different
ways, one of which is to check the weight of powder applied per centimeter
tube length.
For production control optical means may be used, for instance, by mea-
suring the diffuse reflection of a narrow beam of light passing through the
glass envelope onto the coating (see Fig. 5.11).
From a theoretical point of view the optical control is more correct, be-
cause in general the optimum efficiency is obtained, for each grain size, at
the same reflection (Fig. 5.12a) whereas the powder weight at which optimum
efficiency is reached strongly depends on the grain size (Fig. 5.12b).
This may be elucidated by means of a simplified model.
Consider a phosphor consisting of equi-sized spherical particles, which is
applied to the glass bulb in successive layers. The particles are supposed to
be entirely opaque for 254 nm radiation, so that all radiation striking a
64 FLUORESCENT LAMPS

Incandescent
light source

/
Photocell

Baffle
plates

-----------------,--------
Bulb "- Phosphor
coatmg

Fig. 5.11. Principle of a device for measuring the optimum coating thickness using the
diffuse reflection of the coating.

particle will be completely absorbed. Visible light, on the other hand, is only
very slightly absorbed, but strongly scattered. The first layer of particles,
when closely packed, will already cover over 90% of the glass. The second
layer will be deposited preferentially in the pockets that occur between any
three neighbouring particles, closing half of the openings completely and the
others partly. In this way the degree of coverage, and thus the ultra-violet
absorption, is increased to some 97 %, which comes very close to the con-
dition in which the optimum light output of the lamp is obtained. Applying
further layers cannot improve the absorption of the ultra-violet materially,
and because of the increased absorption of visible radiation the light output
of the lamp will begin to decrease. The maximum in the light output of the

~
~
)(
100
dm=Bj..lm
e
'0'
dm=8J..1m
.::! ~~
.....
~ ;;:: ~
Ill Ill 5J.im
:;,
0
.!;:
~ 5 ~
E I.J..Im -~

~
.::! 90 .::190
,.,
CLI
:0:
~ CLI
-~
.E .!2
~ 80 &so
1
2J..1m
l 2J..1m
70 70
~ ~
30 50 70 10 20 30 1.0 50
arbitr. units arbitr. units
- Reflection - Powder weight

Fig. 5.12. Relative luminous flux as a function of (a) diffuse reflection; (b) powder weight
per em tube length with the mean phosphor particle diameter dm as a parameter (schematic-
ally).
LAMP DESIGN AND LAMP MANUFACTURE 65
lamp therefore occurs at about two layers of particles. This applies quite
generally regardless of the size of the particles. It implies that the phosphor
weight necessary to get this optimum degree of coverage must be directly
proportional to particle size. On the other hand, in the range of particle
sizes that occur with phosphors, the scattering power for visible light per
unit weight is inversely proportional to the particle size. Therefore the scat-
tering power of the coating (as measured in the reflection or transmission
of visible light) that gives the maximum light output is independent from
the particle size of the phosphor.
t In the general case of a phosphor with particles of various sizes and
shapes the simple model just described will no longer apply. That the optical
properties are the most satisfactory means for controlling the coating thickness
can also be proven in this case by applying the Kubelka-Munk formulas for
the optical properties of powdered materials on the phosphor coating which
can be considered as an optical diffusing surface. According to these for-
mulas the reflected radiation part R is given by:

f3 sinh Kt . ;(- jt- -) i


R =
(fl + {3) sinh Kt + K cosh Kt
Wlth K = f3
l
I -
f3
+1 - I

where f3 = scattering coefficient, fl = absorption coefficient of phosphor


and t = coating thickness.
Now for a given phosphor f3 depends on the grain size only and is pro-
portional with the surface S of the cross-section of the particles per unit
volume and therefore inversely proportional to the mean diameter dm of
the particles:
(5.1)

ft is a material property of the phosphor and varies from about in the


ultra-violet region to about zero in the visible part of the spectrum.

In the visible region ~-+ 0


f3
1 1
In that case lim - = 1+- (5.2)
" R
-->-0
f3t
(3

Now the optimum coating thickness topt will be found at a certain value
S of the total cross-section of the particles per unit volume. The required
thickness to realise this surface S is therefore proportional to dm. Thus
(5.3)
From equations (5.1) and (5.3) it follows that in the situation of optimum
coating thickness f3t = c 1 c2 = constant and according to equation (5.2) this
will lead to a certain value of R, at which optimum light output will be found
independent of the grain size. t
66 FLUORESCENT LAMPS

Fig. 5.13. Cross-section of the phosphor layer in a fluorescent lamp. Magnification approx.
400x. (a) coarse powder (6% < 6 [J.m); (b) standard powder (40% < 6 [J.m).

As an example Fig. 5.13 gives two phosphor coatings of lamps with op-
timum coating thickness. Fig. 5.13a shows coarse powder, b standard pow-
der. The holes between the grains are clearly visible in both layers, but are
larger in the coarse phosphor. Both layers contain on average the same
amount of grains.

5.8 Processing in the lehr


The object of lehring is to burn out the binder from the phosphor coating
and to outgas the bulb and phosphor. To perform this the bulbs with the
coating are heated up during a certain time. Most types of binders ask for
a relatively high lehring temperature, which is not far below the softening
point of the glass. Lehring time is important too. For optimum lehring one
has to find a good compromise between the completeness of the burning out
of the binder and harmful effects which may set in. For instance, during
lehring sodium is released from the glass, a process which increases with
increasing temperature. This sodium may produce a long-term detrimental
effect on the light output. In some cases the combustion products of the
burning gas can react with the phosphor layer. Although not fully under-
stood, the atmospheric conditions can also influence this process a great
deal, resulting in serious short term drops of light output.
A good lehring leads to a clean outgassed phosphor surface, which can
then act as a sponge and reabsorb contaminants from the atmosphere,
LAMP DESIGN AND LAMP MANUFACTURE 67
storage time between lehring and exhausting should therefore be kept to a
minimum. To burn out the binder sufficient oxygen should be present in the
bulb. In many cases the air in the bulb contains enough oxygen. In other
cases, depending on the binder type and coating thickness air has to be blown
through the bulb during lehring. In gas operated lehrs contamination by
combustion products can easily occur, and ventilation has to be performed
very carefully.
Two important parameters of the lehring process are time and tempera-
ture. Although the best evaluation of optimum lehring is by lamp perform-
ance some less time-consuming means can be of great help in evaluating a
certain selected process. For instance quantitative measurements of the
reflection and emission properties of the phosphor scraped out from the
lehred bulb can give good indications about the progress of the process,
because good correlations exist between lamp performance and these prop-
erties, other things being constant (see Fig. 5.14).

v
/
~
- ........

I
I
I
50
I
1.00 500 600 700
Peak bulb temperature
- during /ehring ("C)
Fig. 5.14. Reflectance of phosphor scraped out from lehred bulbs as a function of peak
temperature during lehring for a given coating. For this typical curve total lehring time
was 2 min. The curve shows a maximum at 600 C and also the detrimental influence of
under- and over-lehring.

In addition analyses of sodium and carbon content may give information


on both over-lehring (too high temperature or too long) and under-lehring
respectively.

5.9 Processing on the exhaust machine


The object of the processing on the exhaust machine is to remove all
undesirable gases and volatile impurities from the lamp atmosphere and
inner lamp parts, to activate the electrodes, to introduce the required amount
of mercury and filling gas and to seal off the lamp. The exhaust operation
and the gaseous content of a fluorescent lamp are critical factors in lamp
68 FLUORESCENT LAMPS

quality, especially as regards depreciation. There are several ways to remove


gases out of the bulb:

I. Evacuation with the proper type of vacuum pump. In the pumping range,
where the mean free path of the gas molecules is less than the diameter
of the exhaust tube, pumping can be done very rapidly. This covers the
pressures down to I torr and is a matter of seconds. At lower pressures
we come in the molecular pumping range where pumping becomes very
slow and takes much time.

2. In order to exhaust a bulb in any reasonable time, a gas or vapour flush


must be used after I torr is reached, which in effect raises the pressure
back into the viscous flow range, where fast pumping is possible. In ver-
tical exhaust machines this flush very commonly is a mercury flush. A
droplet of mercury falls onto the hot bottom end of the bulb, vaporises
very quickly and produces a gas piston of mercury vapour from the
bottom of the bulb moving upwards to the upper part of the bulb. The
undesirable gases, compressed by this piston to a pressure of a few torr
are pushed out through the stem orifice and can be pumped away very
quickly.

3. In horizontal exhaust machines the bulb can be exhausted from both sides,
because each of the two mounts is provided with an exhaust tube. After
a rough evacuation to about I torr in a few seconds, an argon flush
through one exhaust tube is applied, at the same time exhausting the bulb
through the other exhaust tube. As contrasted with the situation under
1 and 2 the pressure does not fall below say 1 torr, because the equili-
brium pressure of the flushing gas is well above this value. Therefore on
this exhaust machine no costly pumps with very low end vacuum (e.g.
diffusion pumps) are necessary. By using for the flushing gas the same
rare gas as is used for the filling gas pumping down to the required filling
pressure is sufficient. A droplet of mercury can be blown into the bulb by
means of a small fixed quantity of the rare gas.

During the operations on the exhaust machine the temperature of the


lamp has to be relatively high, the desirable level depending upon the specific
aspects of the machine group. The reasons are:

l. Increase of the molecular motion of the gases and thus enhancing at low
pressures the probability of a gas molecule striking the stem orifice.

2. Gas molecules trapped in or attached to the glass, phosphor and lamp


parts are released more readily when heated.

3. Moisture and other condensed vapours are volatilised by heat and


removed more easily.
LAMP DESIGN AND LAMP MANUFACTURE 69
5.10 Activation of the electrodes
A very important feature of the processing on the exhaust machine is the
activation of the electrodes. As has been pointed out earlier in this chapter
the electrodes are designed for a low work function and long cathode life.
The most satisfactory jiolution so far has been a tungsten coil covered with
an oxide coating which consists of a mixture of barium oxide with other
earth alkaline oxides (Ca, Sr). In order to get a low work function the mixed
oxide must contain a small amount of free barium.
The tungsten coil (coiled coil or triple coil) is coated with an intimate
mixture of the earth alkaline carbonates. In the activation process the elec-
trode is heated so as to drive off the remainder of the binder, which has
already been partially decomposed by the heat of the sealing fires and the
subsequent exhaust oven heat. At still higher temperatures realised by an
electrical current through the coil, the carbonates are converted into oxides,
which react with tungsten metal to give the required amount of metallic ba-
rium. To get an emitter which can withstand the eroding action of the gas
discharge some additives to the earth alkaline carbonates may be applied,
for instance Zr, AI or their oxides.
The activation schedule depends on the pumping scheme, especially in
relation to the speed with which the developed carbon dioxide can be pumped
or flushed away without any tendency to concentrate at the electrode. Norm-
ally the temperature in the beginning of the schedule is relatively low, but
during the last part of the degassing period it is necessary to increase the
temperature for a short time, although overactivation has to be prevented.
If the temperature rise in the beginning of the activation is too sharp, the
bond between the tungsten and the emission material may be ruptured,
resulting in subsequent flaking of the cathode coating. In practice it is found
that with optimum cathode conditions a slight amount ("' 1 %) of C0 2
still remains in the emission material. It seems that a complete degassing
of the cathode drives off too much of the barium.
70 FLUORESCENT LAMPS

Fig. 5.15. A modern fluorescent lamp factory.


Chapter 6

Colour and colour rendering


A. A. Kruithof f j. L. Ouweltjes

6.1 Blending of fluorescent materials


In Chapter 4 we saw that fluorescent materials as employed in fluorescent
lamps are capable of emitting light of widely divergent colours. Calcium
silicate, for instance, when activated with lead and manganese, gives an
orange; zinc silicate, activated with manganese (willemite) a green, and cal-
cium tungstate a blue-violet emission. Some of the colours are very vivid
because the spectral emission bands of many of these materials are rather

Fig. 6.1. When yellow and blue paints are mixed, the resultant colour is green ; the same
result is obtained when yellow ink is printed on top of blue ink.
72 FLUORESCENT LAMPS

---------

/I
I I
I
I

M~---1--~IL--hti-
/ I
j I i
j I \
\
I I \
~'~--~~~-~~+ \
'---+-----
-------- /
/
I

... ..-,/
600 700
---A.(nm)

Fig. 6.2. Mixing of yellow and blue pamts. e = reflection factor. Full line = curve for
yellow paint. Broken line = curve for blue paint.
Dot-dashed line = curve for mixture. A = wavelength in nm. The colour impressions
evoked by the various wavelengths are indicated at the top of the diagram.

narrow. If the light from a fluorescent lamp is required to cover the whole
of the visible spectrum, often two or more different materials must be mixed.
When fluorescent materials are blended the results are by no means the
same as with paints. It is well known, for example, that yellow and blue
paints when mixed produce a green, as shown in Fig. 6.1. The reason for
this is illustrated in Fig. 6.2 which shows how parts of the spectrum are
reflected by the two paints; the rest of the light is retained, i.e. absorbed.
The yellow paint absorbs the blue light; the blue paint absorbs the yellow
and the red. When the paints are mixed they are diluted and thus each of
them absorbs less light. However, the incident rays of light now encounter
both yellow and blue particles in the layer of paint. The resultant reflection
curve, the dot-dashed curve of Fig. 6.2 shows a minimum in the blue spectral
region which is due to the yellow particles and in the yellow another mini-
mum due to the blue particles. Between the two, in the green, a maximum
occurs since both paints are fairly reflective in that part of the spectrum.
The result is very different, however, when a mixture of a yellow- and a blue-
emitting fluorescent material is applied to the inside wall of a fluorescent
lamp. Spectral energy distribution curves for two such substances are given
in Fig. 6.3. One of the materials emits mainly yellow, orange and red radia-
tions, i.e. yellow light, whilst the other emits blue, green and violet, i.e. blue
light. By adding increasing proportions of the blue phosphor to the yellow
one, the latter is successively more diluted so that the yellow emission de-
creases and at the same time the blue emission of the mixture increases. In
COLOUR AND COLOUR RENDERING 73
100 woe
.f t blue green yeflow orange
,
I
1/~ I I/ l'\
1\ I
I
I I
I
\

,I \
0
I I
I I
I \
I
I I

----......-,
I

I1/
I
I

~/
I

,/-- ......... ....... -------


/
I
\ I
I \ \
/ I \

\
/ \ \
I \
I
I \ \
I \
I '
I
/
I
I \
\ '\
I
-,.,
I
20

~
I \
I \
I ' .,_\ ',

I
I
I
I
r---- ....... ~
-.
500 600 700
- - - - A(nm)

Fig. 6.3. Blending of yellow- and blue-emitting fluorescent substances. E= relative amount
of power emitted at wavelength ..1.. Full line = spectral energy distribution of blue-emitting
substance. Broken line = spectral energy distribution of yellow-emitting substance. Dot-
dashed line = spectral energy distribution of the mixture.

this way a ratio of the yellow- to the blue-emitting material can be chosen
for which the energy radiated by the mixture is found by adding half the
energy given by the yellow phosphor to half that emitted by the blue phos-
phor for every wavelength of the visible spectrum. Fig. 6.3 shows that this
particular blend would give us a lamp emitting a fairly regular spectrum
throughout the whole of the visible range (wavelength 400-750 nm). A lamp
of this kind would therefore emit white light as illustrated in Fig. 6.4.
If the mixture contains a higher proportion of yellow-emitting material
than that assumed for Fig. 6.3, the light is of a more yellowish white; on
the other hand if the blue-fluorescing substance predominates, the light is

Fig. 6.4. When yellow- and blue-emitting fluorescent materials are mixed, the resultant
fluorescent light is white.
74 FLUORESCENT LAMPS

more blue in tint. In this way all intermediate tints between yellow and blue
can be obtained by varying the proportion of the component materials of
the mixture. When three fluorescent substances are mixed together, each
giving respectively red, green and blue light, intermediate tints between those
three extremes are produced in the same way. In order to appreciate the
significance of this statement it is necessary to consider first some of the
factors affecting our own capacity for seeing colours.

6.2 The fundamentals of colour vision


Colour vision enables us to tell whether objects around us are red, orange,
yellow, green, blue or purple, or whether their hues are somewhere between
these main hues. We can also judge whether colours are vivid or weak, light
or dark.
Researches by Young * as far back as the first years of the last century
prompted Helmholtz to put forward the hypothesis that the presence in the
eye of three kinds of photosensitive receptors enables us to distinguish
colours, a hypothesis that has recently been confirmed by experiment. We
are indebted to Grassmann for three laws allowing us to assume that the
results of the processes within the eye can be represented by the currents in
three photo-electric cells, each cell having its own spectral sensitivity-curve.
These laws have been checked by numerous experiments.
One such photo-electric cell is sensitive mainly to red light, the relative
sensitivity being represented by curve x in Fig. 6.5. The second cell, y, is
sensitive mainly to green light, and the third, z, to blue. The sensitivities of
these cells need not necessarily correspond to those of the receptors in the
eye. In fact the latter sensitivities are very difficult to determine with any
degree of accuracy. But the sensitivities of the photo-electric cells are so

Fig. 6.5. Sensitivity curves of


the three photo-electric cells
imitating the function of the
human eye (C.I.E. 1931 stan-
dard colorimetric system).
G = sensitivity
:X = sensitivity of the 'red'
cell.
y = sensitivity of the 'green'
cell.
z = sensitivity of the 'blue'
cell. 700
J. = wavelength in nm.

* The points mentioned in Sections 6.2, 6.3 and 6.4 are dealt with more fully in a work
by Dr. P. J. Bouma 'Physical Aspects of Colour'.
COLOUR AND COLOUR RENDERING
Table 6.1
C.l.E. 1931 standard system of colorimetry.
Sensitivity values of the 'red' photo-electric cell X, the 'green' cell Y and the 'blue' cell
Z. A = wavelength in nm.

A :X y
I -z
I A
I -
X
-y
z
380
390
0001
0004
0
0
0006
0020
600
610
620
I 1062
1003
0854
0631
0503
0381
0001
0
0
400 0014 0 0068 630 0642 0265 0
410 0044 0001 0207 640 0448 0175 0
420 0134 0004 0646
430 0284 0012 1386 650 0284 0107 0
440 0348 0023 1747 660 0165 0061 0
670 0087 0032 0
450 0336 0038 1772 680 0047 0017 0
460 0291 0060 1669 690 0023 0008 0
470 0195 0091 1288
480 0096 0139 0813 700 0011 0004 0
490 0032 0208 0465 710 0006 0002 0
720 0003 0001 0
500 0005 0323 0272 730 0001 0 0
510 0009 0503 0158 740 0001 0 0
520 0063 0710 0078
530 0166 0862 0042 750 0 0 0
540 0290 0954 0020 760 0 0 0

550
560
570
0433
0594
0762
0995
0995
0952
0009
0004
0002
I 1068
I 1068
I 1068

580 0916 0870 0002


590 1026 0757 0001

closely related to those of the receptors that they can be used to represent
the function of the eye. This relation may be found from experiments in
which the spectral colours are matched by mixing the light of three fixed
primary light sources (radiating for instance a red, a green and a blue spec-
trum line respectively). It was on the basis of very careful measurements of
this kind, made by Guild and Wright, that the C.I.E. *in 1931 standardised
the three curves shown in Fig. 6.5. The ji curve was chosen by the C.I.E.
to be identical with the visibility curve V(A.). Table 6.1 gives the values of
the sensitivities for 10 nm wavelength intervals in greater detail than is
possible in Fig. 6.5. The three curves x, ji and z enable us to construct a
diagram within which each of the colours the eye can perceive is represented
by a point, thus producing as it were a map of the whole colour domain.
This map is called the chromaticity diagram.
* International Commission on Illumination. (French: Commission International de
l'Eclairage.)
76 FLUORESCENT LAMPS

6.3 Construction of the chromaticity diagram


The construction of a two-dimensional chromaticity diagram representing
all colours is possible because of the fact that, within certain limits, the
amount of light emitted by a coloured surface makes very little difference to
the colour impression e.g. a piece of red paper has the same colour whether
it is 1-, I or 3 m from the same light source. The proportions of the amounts
of light falling on the paper at those distances would be as 36 : 9 : I.
For the three photo-electric cells analoguous to the eye, halving the
amount of light falling upon the piece of paper means halving the photo-
current of each cell. The chromaticity must not then be changed; in other
words, for judging colours only the ratios of the photocurrents are impor-
tant.
It is easy enough to express colours in numbers if the cells x, y and z
-as shown in Fig. 6.5 - are available. All that is then necessary is to allow
the light of which the colour is to be characterised to fall onto the three cells,
measure the three photo-currents, X, Y and Z, and compute the following
relations
X y z
X=----- Z=-----
X+Y+Z y=X+Y+Z X+Y+Z

Any change in X, Y and Z by the same factor in each of them will thus
leave the values of x, y and z unchanged. The chromaticity co-ordinates x,
y and z are therefore independent of the amount of light and, as such, are
suitable for representing the colour. Any pair of these taken together, such
as x andy for example, is sufficient to determine the chromaticity. Since
X+Y+Z
X+ y +Z = = 1,
X+Y+Z
z can be evaluated when x and y are known; colour can thus be represented
in a plane figure on a system of co-ordinates x andy. The points in the dia-
gram are then termed colour points. As in Fig. 6.6, the x co-ordinates are
generally plotted horizontally and the y co-ordinates vertically. Seeing that
photo-electric currents are always positive or zero, X ~ 0, Y ~ 0 and
Z ~ 0; then x ~ 0, y ~ 0 and z ~ 0, so that, where x + y + z = 1, it
can be said that x ~ 1 ; y ~ 1 and even that x + y ~ 1. In consequence,
the whole chromaticity diagram is enclosed within a triangle bounded by
the x- and y-axes and the line x + y = 1.
The simplest way to determine the colour point of any coloured light is,
already pointed out, to use three photo-electric cells of the appropriate
spectral sensitivities. There are, in fact, many instruments that work along
these lines, and these are called physical colorimeters. Unfortunately, how-
ever, it has not been found possible by simple means, such as by the use of
glass filters, to make the sensivity curves of photo-electric cells exactly equal
to the curves shown in Fig. 6.5. The only satisfactory precision method
known up to the present for the direct determination of colour co-ordinates
is to split up the coloured beam of light into its spectral components with
COLOUR AND COLOUR RENDERING 77

0,4 0.6 0,8 1,0


-x
Fig. 6.6. Chromaticity diagram plotted on x and y co-ordinates. All colour points lie
below the line x + y = 1. The diagram includes the colour points of the spectral colours
of which the wavelengths are indicated. Further the 'white point' W is indicated. with
co-ordinates x = t, y = t and the purple line p. Along the straight line joining W to
the colour-point of the spectral colour A.= 470 nm, blue colours of increasing paleness
are found the nearer W is approached.

the aid of a (single) prism monochromator. A template is placed in the


spectrum. After having passed the template the light is sent through an
identical monochromator in the reverse direction. The spectral components
are then brought together once again on the slit, and fall on a photo-electric
cell. The template is used to reduce radiations in spectral regions where the
combination of double monochromator and cell is relatively more sensitive
than it should be according to Table 6.1, say for the .X photo-electric cell.
For the ji and z cells other templates are used. In this way the C.I.E. sen-
sitivity curves can be imitated with good precision.

6.4 Computing colour points


The colour co-ordinates of a coloured light may be calculated without
further measurement if the spectral energy distribution of the light is known.
It is very simple to determine the chromaticities of spectral or mono-
chromatic light sources. For example, from Table 6.1 the following values
78 FLUORESCENT LAMPS

are obtained for light having a wavelength of, say, 480 nm; X= 0096,
Y = 0139, Z = 0813, so that x = 0092 andy= 0133.
When these calculations are carried out for the spectral colours mentioned
in Table 6.1, the points shown in Fig. 6.6 are obtained. A smooth curve
through these points contains the colour points of all spectrum colours and
is called the spectrum locus. For a mixture of light of two or more spectral
colours the photo-electric currents are obtained by adding the corresponding
currents of the components. It should be noted here that - since the y curve
is identical to the standard visibility curve V(A.) * - instead of the Y values
of the components, the lumen values can be taken.
Whilst the colour points of monochromatic light sources and mixtures of
light from two or more such sources are easily determined, even without the
aid of the three photo-electric cells imitating the human eye, it seems more
difficult to do so for the light of the incandescent-filament lamp, for natural
daylight, or for the emission from fluorescent lamps. Such light is of a much
more complex nature: all wavelengths of the visible spectrum are emitted
more or less strongly. Measurement of the spectral energy distribution of an
incandescent lamp gives a curve as the one of Fig. 6.7, which gives the rela-
tive amount of energy emitted at every wavelength.

violet blue green yellow orange red


250
E I I I I I
/
/
200

150 /
/
/
50
v
/
v
0
400 ------- 600
-.ArnmJ
?00

Fig. 6.7. Spectral energy distribution curve of the incandescent filament lamp (2 854 K).
E = relative radiant power emitted at wavelength A..

Let us now see how one can determine the colour point of such a light
source from the spectral energy distribution curve using the sensitivity curves
of the photo-electric cells given in Table 6.1. It is first assumed that each
sensitivity value given represents the average sensitivity of the photo-electric
cell within a narrow band of wavelengths, say 10 nm in width, although the
* According to the C.I.E. vocabulary this curve is named the spectral luminous efficiency
curve for the standard photometric observer.
COLOUR AND COLOUR RENDERING 79
sensitivity of the cell of course varies within such a band. For the point
A = 530 nm for example the band then would extend from 525 to 535 nm.
The average radiant energy of the light source, within a zone 10 nm in width
about the point A = 530 nm, can be found by reading from the spectral
distribution curve in Fig. 6.7 the relative amount of energy E 530 corres-
ponding to that wavelength. The energy emitted in the zone will then be
the product of the average energy and the width of theb and 530 x 10.
The contribution of the 10 nm range of the photocurrent of the 'green'
photo-electric cell for example is again the product of average sensitivity
and radiant energy, that is ji 530 X E 530 x 10. Similarly, the components of
the photo-current for each of the 37 wavelength zones of 10 nm lying be-
tween 375 nm and 745 nm are ascertained*. The total Y photo-current is
the sum of the 37 components. Denoting summation by ~ (sigma), the Y-
current can be represented by

Similarly, X= ~ x;.E;. X 10, and Z = ~ i;.E;. X 10.


Once X, Y and Z have been evaluated, the colour point is easily located
in the manner described in Section 6.3.
The calculation becomes very simple indeed if E is the same at all wave-
lengths and therefore in all terms of the above summations (equal energy
spectrum).
It is then found that

X= ~X;. X Ex 10 = IOE ~X;., i.e. x =

Y = ~ ji;.XEX 10 = IOE ~ )i;.

Z = ~ i;. X Ex 10 = 10E ~ i;.

The C.I.E. have so standardised the sensitivities of the photo-electric cells


that ~X;.= ~ )i;. = ~ i;. which at once appears from the summation of
the vertical columns in Table 6.1. From this it follows that there x= y = t,
that is to say the centre of gravity of the colour triangle (Win Fig. 6.6) is
the colour point of the equal energy spectrum. This equal energy spectrum
closely resembles that of daylight, and the impression on the eye is white.
Thus the centre of gravity of the colour triangle is known as the 'white'
point.
The spectral energy distribution of a blended colour composed of light
from two arbitrary coloured light sources can be found by a method analo-
* According to the C.I.E. the visible range of the spectrum extends from 380 nm to
760 nm. The values of x, y and z were therefore given for 39 wavelength points and the
summations should be made for 39 bands. In Table 6.1, however, the values are given
to only three decimal places and in this approximation for the wavelengths 750 nm and
760 nm they are all zero.
80 FLUORESCENT LAMPS

a b c
Fig. 6.8. A mixture of yellow and blue light gives the same result as a mixture of fluores-
cent substances. The figure shows how a white surface appears when illuminated with
(a) yellow light, (b) blue light, (c) both colours simultaneously.

guous to that described for the colour of a mixture of two fluorescent sub-
stances in Section 6.1. At each wavelength the energy emitted in the 'mixed'
light is the sum of the energies of the original components. Such mixed light
for instance originates by projecting two beams of light together on a white
sheet in the manner shown in Fig. 6.8. We shall now derive the colour point
of a blend of two coloured beams. Let P 1 be the colour point of the first of
the two light beams (see Fig. 6.9), giving photo-currents Xt. Y1 and Zt.
and P 2 that of the second light beam giving currents X 2 , Y2 and Z 2 ; the
photo-currents Xm , Ym and Zm of the mixture are then
Xm = X 1 + X2 ; Ym = Y1 + Y2 ; Zm = Z 1 +Z 2

xl yl
Now x 1 = - - - - - - Yl= - - - -- -
Xl + Y1 + zl Xl + Y1 + zl
Yz
Xz = - - - - - - Y z = - - -- - -
Xz + Yz + Zz Xz + Yz + Zz
-- - ------ --- ----- ~

YrYt
-- - -- -- - - 71

P,
COLOUR AND COLOUR RENDERING 81
whilst for the mixed light
X1 +X2
Xm=--------------------------
+ X2 + Y1 + Y2 + Z1 + Z2
X1
y1 + y2
Ym =--------------------------
X1 + X2 + Y1 + Y2 + Z2 + Z1
It is easily deduced from these relations that

Ym- Y1 Y2- Y1
and
y1 y2
(Y2- Ym): (ym- Y1) = - : -
Y1 Y2

The first of these relations means that the triangles BP 1 M and AP 1P 2 in


Fig. 6.9 are similar, so that M must lie on P 1P 2 The second indicates that
M divides the distance P 1 P 2 in the ratio of Ydy 1 : Y 2 /J2. The values Y1
and Y 2 are proportional to the quantities of light in the two original beams,
and we can therefore say that as the mixture of light of two colours, repre-
sented by points P 1 and P 2 , contains more of the colour P 1 , the colour point
will lie nearer to P 1 on the straight line P 1P 2
This feature can be immediately applied to determine the colour points
for mixtures of spectral colours and especially for blends of the deep reds
and far violets found at the two ends of the spectrum. The colours of those
blends, of which the proportion of the components may be variable, are
purple, and the colour points lie on the oblique line p in Fig. 6.6, which
joins the two ends of the spectrum.
As stated in the foregoing, the colour points of the spectral, i.e. vividly
coloured types of light are situated in the spectrum locus in Fig. 6.6. The
rule for mixed light now gives us further information about the other points
in the triangle. If a line is drawn from any point on the spectrum locus or
on the 'purple' line (for example the point kin the figure where A. = 480 m)
to the 'white' point, it will be seen that any point of the colour triangle can
be obtained by mixing spectral or purple light with white light. The more
white is introduced into the mixture, the more the spectral colour becomes
diluted, or, as we say, the less saturated. The colour point of the mixture
is then shifted more towards the white point W.
The 'colour map' is reproduced as well as printers ink will allow, in Fig.
6.10. It is obvious that the true spectral colours are much more vivid than
the most highly coloured printing inks, but the figure at any rate provides
an impression of the locations of the colours in the chromaticity diagram.
82 FLUORESCENT LAMPS

6.5 Application of chromatics in the development of fluorescent


lamps

The principles concerning the colour of light and light sources that have
been evolved in the preceding sections can now be applied in deciding what
colours of light may be obtained from fluorescent lamps. The light emitted
by these lamps consists for the greater part - 88 to 95% - of the fluores-
cent light from the substance or mixture of substances with which the tube
wall is coated. The rest of the light is generated directly by the mercury
discharge. The mercury spectrum contains a number of spectrum lines, of
which the principal ones occur at the wavelengths A = 405 nm, 408 nm,
436 nm, 546 nm and 578 nm. This, however, presents no obstacle in the
calculation of the colour points; it is merely necessary to add the contribu-
tions XHg YHg ZHg coming from the spectral lines to the photo-currents
XI> Y, and Zf, produced by the fluorescent light. Measurements to ascer-
tain the spectral distribution of the fluorescent light give us at the same time
the intensities of the mercury lines, and these in turn give the correction to
be made in the chromaticity to allow for the 5-12% of mercury light mixed
with the fluorescent light.
The results of measurements on lamps separately coated with each of the
individual phosphors, commonly applied in fluorescent lamps, are used to
calculate the colour points. The resulting points which, owing to the cor-
rection for the presence of the mercury light are somewhat more to the left
than those for the pure fluorescent light, are reproduced in Fig. 6.11 for the
lamps containing calcium silicate, willemite and calcium tungstate. Com-
parison with Fig. 6.10 shows that the colours of the light emitted by the
lamps are as follows: calcium silicate (point C.S.)- red; willemite (point
Wil) - green; calcium tungstate (point C.T.) - blue-violet.
For the halophosphates which, according to Chapter 4, can be made in
a great variety of colours, the colour points are found near the white point W
in Fig. 6.11 with x varying roughly from 0 3 to 0 5 and y from 0 3 to 04.
An example is given in point H. Finally we have indicated the phosphor
magnesium arsenate (point M.A.).
As the visibility curve V(J.) is the same as the y curve of Fig. 6.5, it is
obvious that phosphors having their emission band in one of the extremes
of the visible range (blue or red) give much less light than those emitting
in the centre of that range although the energy of the emission may be the
same or even sometimes more for the first-mentioned phosphors.
It has already been explained in Section 6.1 that the blending of two phos-
phors amounts to the same thing as the mixing of two different kinds of
light, as far as the spectral distribution is concerned. After the chromaticity
of the mixture has been found, the correction for the mercury lines is applied.
As long as the phosphors do not absorb visible light, the corrected chromati-
city of the mixture is also obtained by applying the mixing rule, discussed in
Section 6.4, to the corrected colour points of lamps which have been coated
with the individual phosphors.
From the mixing rules it follows that blends of willemite with calcium
silicate give all the tints the colour points of which lie on a straight line
COLOUR AND COLOUR RENDERING 83
1.0

111---50-0---+----=-=ir---~

0 0.6
-----lX
0.8 1.0
Fig. 6.11. Colour points of lamps coated with the following fluorescent materials,
C.S. calcium silicate,
Wil. willemite,
C.T. calcium tungstate,
H. halophosphate,
M.A. magnesium arsenate,
I a mixture of calcium silicate and willemite
II a mixture of calcium silicate, willemite and calcium tungstate,
W 'white point'.

drawn from Wil. to C.S. in Fig. 6.11. A mixture of a certain proportion


produces a tint corresponding, say, to the point /. If calcium tungstate is
then added, the colour point of the mixture of the three substances lies on
the line 1-C.T., say at II. As there is complete freedom of choice in the
ratios of the constituents, so that I may be taken to be anywhere on the
line Wil.-C.T., the line 1-C.T. can traverse the whole triangle C.T.-Wil.-C.S.
The point II can similarly occur anywhere on the line 1-C.T. So, as a con-
sequence of the freedom of choice for the two mixing ratios the chromaticity
of the mixture of three phosphors may occur at any point within the triangle
of their colour points. On the other hand closer consideration shows that
a given colour point within the triangle can always be obtained with only one
mixture of the lights of three given components.
84 FLUORESCENT LAMPS

Further if we start with a halophosphate having the colour point H, it


may be corrected to give the white point W, by addition of some calcium
tungstate phosphor.
A glance at Fig. 6.10 shows that, as a result of the mixing rules for the
colour of the light emitted by fluorescent lamps, these lamps can be made
in pronounced red, yellow, green or blue tints, or on the other hand in white,
pale red, yellow, green, blue or pink. Now which shades are suitable in
practice?
Either a vivid or a soft colour may be suitable of course for decorative
illumination. In order to produce a more saturated colour, lamps containing
a red phosphor - e.g. magnesium arsenate or calcium silicate - have been
covered with a layer of red lacquer, or a red pigment is applied to the inside
of the bulb. Thus the violet (436 nm) and the green (546 nm) mercury lines
are filtered out and the colour of the lamp becomes a pure red. The same
procedure has been followed for yellow lamps, where the pigment is yellow.
For blue and green lamps it is not necessary to use pigments. In the following
we propose to limit our review to those tints which are suitable for general
lighting purposes.

6.6 Choice of colours for general lighting fluorescent lamps


In the early 1930s the first attempts were made to develop a lamp for
general lighting purposes from the low pressure mercury discharge by coating
the inner glass wall of the tube with fluorescent powder. The only rule then
available for the choice of the colour. was that the lamp should have a 'white'
appearance. Further investigations showed that 'white' is by no means a
well defined colour. The qualification 'white' is given to a tablecloth in day-
light as well as in incandescent lamp light and even in candle light, although
in all these cases the eye receives completely different radiations from the
cloth. This is an example of chromatic adaptation.
We are only aware of the differences between the various white radiations
if they are present at the same time in the visual field. This situation is en-
countered rather often in fluorescent lighting. For example the lamps may
serve as additional lighting in dark shops and offices or they may be used
in combination with incandescent lamps that are necessary for certain
directional effects. When large colour differences exist between combined
light sources, coloured fringes will occur at the boundaries of all shadows.
These annoying phenomena had of course been noticed before the intro-
duction of fluorescent lamps when daylight and incandescent lamp light had
sometimes accidentally been combined. At that time the fringes had been
accepted as generally unavoidable. After the introduction of fluorescent
lamps, however, when the range of available phosphors permitted a large
freedom in the choice of lamp colours, it was no longer necessary to tolerate
coloured fringes and thus the above vague rule had to be elaborated.
Now filament lamps, the sun and even candles have this much in common
that the light is radiated by incandescent particles or bodies.
COLOUR AND COLOUR RENDERING 85
According to well-known physical laws as formulated by Kirchhoff, Wien
and Planck, the spectral energy distribution of an incandescent body at a
certain temperature is independent of the nature of the body, provided that
it possesses the characteristic to reflect none of the rays falling upon it from
the outside. In this special case the spectral energy distribution of the radia-
tion can be computed and the chromaticity readily determined along the
lines laid down in Section 6.4. A body that reflects none of the incident
radiations is called a 'black' body or a perfect radiator. The former name is
seemingly inappropriate in so far as a 'black' body at the temperature of the
sun is anything but black, emitting as it does a blinding white light. The
temperature of an incandescent black body is usually given in degrees Kelvin
which begins at absolute zero, so that the melting point of ice is 273 K and
the boiling point of water 373 K. Fig. 6.12 shows a number of colour points
of the perfect radiator. The curve through these points is named the black-
body locus and it gives the chromaticities at intermediate temperatures as
well. It has been found by actual measurement of the spectral energy dis-

1.0

lD
__ ___,~X

Fig. 6.12. Chromaticity diagram with black body locus showing chromaticities of the full
radiator. The numbers give the temperatures in thousands of degrees on the Kelvin scale.
Daylight= colour point of 'Daylight' ('Colour matching') fluorescent lamp: Tc = 6 570 K
C. White= colour point of 'White' ('Daylight' fluorescent lamp) Tc = 4300 K
W. White= colour point of 'Warm White' fluorescent lamp: Tc = 2900 K. The colour
names given in brackets are those used in England.
86 FLUORESCENT LAMPS

tributions that incandescent solids, such as metal filaments, the sun, or candle
flames are not exactly perfect radiators but that they nevertheless often have
the same colour as an incandescent black body, provided that the tempera-
ture of the latter be appropriately chosen. This temperature, which can be
either higher or lower than that of the incandescent solid, is called the colour
temperature (Tc) of the light source. The colour temperatures of a series of
well-known light sources are given in Table 6.2.

Table 6.2

Colour temperatures of some natural light sources

Light source Colour temperature in K

Blue sky 10 000-20 000


Overcast sky 5 000-7 500
Direct sunlight ca 5 000
Carbon arc 4000
Photographic lamp 3 200
Incandescent lamp for general lighting 2 400-3 000
Candle 1 900

Taking into account the fact that the usual light sources for interior light-
ing have the same colours as perfect radiators, of which the colour points
appear on the black body locus shown in Fig. 6.12, we can now investigate
the consequence of a more stringent rule to which the colours of fluorescent
lamps must conform. In themselves, and by reason of the contrasts which
they may produce, the lamps must neither have an unpleasant colour, nor
must they make other light sources in the same surroundings look unpleasant.
One fact that is known from experience, and which must certainly be allowed
for when applying this rule, is that light sources having a green or a purple
tint are more annoying than those which appear to be yellowish or blueish
in character. From the chromaticity diagram in Fig. 6.10 it is seen that a
yellow-blue contrast may occur between two light sources having different
colour points which both lie on the black body locus above 2 000 K, whilst
an unpleasant red-green contrast will become apparent if the colour point
of one of the sources is at some distance from this line, the other one lying
virtually on the line. The final requirement to be met by the colour points of
the lamps may thus be formulated quite simply by stating that they must lie
on or very close to the incandescent black body locus. Wholly on the basis
of the definition, then, a colour temperature can be attributed to these light
sources. Having regard to this rule, the tints finally selected will be governed
very largely by the purpose for which the lamps are to be employed.
The range of colour temperatures of lamps indicated as 'Daylight' lamps
(in England 'Colour matching' lamps) extends from 6 500 K to 7 500 K.
In order to obtain more uniformity, international discussions have been
COLOUR AND COLOUR RENDERING 87
held in the I.E. C. * to choose one of the colour temperatures of this range
for all lamps which will be indicated with those names.
For the 'Warm White' colour corresponding to incandescent-lamp light
the range of colour temperatures is 2 600 K- 3 100 K.
In addition to lamps which are, if necessary, suitable for combination
either with daylight or with incandescent lamp light, there was a need for
lamps that can be combined with both these kinds of light. In order to be
able to explain the choice of a colour for such a 'Cool White' lamp (in
England 'Daylight' lamp) we shall once more make use of chromatics. It is
clear that a perfect adaptation to both daylight and incandescent lamp light
is impossible so that ultimately a compromise must result. An obvious choice
for such a compromise would be the colour point which is situated midway
on the line connecting the colour points for an average daylight (Tc = 6 500 K)
and incandescent lamp light (Tc = 2 850 K). As a consequence of the cur-
vature of the black body locus this point is situated below that curve. The
colour contrasts with daylight and incandescent lamp light appear slightly
smaller, however, for a point virtually on the black body locus. The most
suitable colour temperature comes out nearer to that of the incandescent
lamp than to that of daylight in the region 4 000-4 500 K.
In addition to the lamps specified in the foregoing, there is a colour
'3 500 White' which has its colour point on the blackbody locus at Tc =
3 500 K. In England there is also one more colour, namely 'Natural',
having its colour point well below the black body locus near the colour
temperature Tc = 4 100 K.

Fig. 6.13. Spectral energy distribution curves for 'Cool Daylight' fluorescent lamp and
full radiator at 6 750 K.
Jc = wavelength in nm.
E/rp = energy flux in to- 6 W for a wavelength band of 10 nm width per lumen light flux.
The energy of the mercury spectrum lines, which in fact is concentrated in extremely narrow
bands, is represented as 20 nm wide bands on top of the spectrum of the phosphor.

* International Electrical Committee.


88 FLUORESCENT LAMPS

--- V-~~~ ~--- ------ -----


r--

U.6 --. .
~
1i .,4300_!<

""I'--
c;_-- ,'
../
~hite
!---'""""' _.,
t---
400 500 600 700
-)\, (nm)

Fig. 6.14. Spectral energy distribution curves for 'White' fluorescent lamp and full radiator
at 4300 K.
A = wavelength in nm.
Efq; = energy flux in to- 6 W for a wavelength band of 10 nm width per lumen light flux.

Examples of the three types of lamps are the so-called standard fluorescent
lamps, which are universally made with halophosphate phosphors, and are
therefore very similar in their properties. To promote interchangeability
more exact chromaticity values have been internationally standardised. The
spectral energy distributions are shown in Figs 6.13, 6.14 and 6.15.

----
300
E!p
........ ........
r -
....

t ........
........

'\
~ 200 \.
~

~
l::l.~
r:l::l. r-- v........--
f ....
.... ....
100

o--
I(
-zgoo............
....
'\
I

_/
I

1c _.,.,

.
-~rm White

sao 600
~r- 700
_ ____,., A. ( nm)
t'
Fig. 6.15. Spectral energy distribution curves for w arm White' fluorescent lamp and full
radiator at 2 900 K.
A = wavelength in nm.
E/q; = energy flux in to- 6 W for a wavelength band of 10 nm width per lumen light flux.
COLOUR AND COLOUR RENDERING 89
6.7 Colour tolerances
When fluorescent lamps are mounted at close distances, large differences
between the tints of adjacent lamps tend to produce discomfort. Such dif-
ferences are especially noticeable if the lamps are mounted end-to-end in
rows. It is therefore important to know which differences in tint will prove
troublesome and which will not.
Laboratory tests can be so arranged as to make the eye extremely sensitive
to colour differences; two tints are presented to the observer at equal bright-
ness levels on adjacent fields, e.g. by means of the circular comparison field
in a colorimeter shown diagramatically in Fig. 6.16. Provided that this field
is sufficiently large, far smaller differences can be detected than in the crude
test when lamps are merely placed in a row. A large number of such labora-
tory tests have been carried out in which light of a certain chromaticity is
thrown on one of the halves of the field. When light of the same chromaticity
is allowed to fall on the other half, the two halves are indistinguishable and
the line of demarcation disappears. A slight change in the light falling on
one of the halves means a shift of the colour point and, when the change is
sufficiently great, the difference in colour between the two halves of the com-
parison field becomes apparent. In this way it is possible to find the distance
between two points representing colours showing a just perceptible difference.

Fig. 6.16. Comparison field of a colorimeter. Ofthe two co-


lours to be compared one illuminates the left-hand side and
the other the right-hand side of the field.

The distance, which depends on the direction of the line joining the colour
points and on the position they occupy in the chromaticity diagram, is named
the 'minimum perceptible colour difference'. It has been found that in the
region of the colour temperature Tc = 4 300 K a temperature difference
between 15 K and 20 Kjust produces a perceptible variation. When a smaller
comparison field is used and also under the practical conditions prevailing
when comparisons between fluorescent lamps are made, the just visible
variation is much larger than the minimum perceptible difference. In the
latter case it is even some 10 times greater. Furthermore, a barely visible
variation will not quite be annoying. It can thus be said that, as long as the
distance between two colour points is smaller than about 20 minimum per-
ceptible differences, the colour difference, although it may be visible, in
practice is not annoying. In directions perpendicular to the black body locus,
however, somewhat greater care must be taken than along that line. The
colour difference represented by a distance of 20 times the minimum per-
ceptible difference will, for the purpose of this book, be called one 'step' in
90 FLUORESCENT LAMPS

the chromaticity diagram. A difference in colour amounting to less than one


step produces thus, for practical purposes, no discomfort. If points are plotted
in all directions about the colour point W of the colour 'White' for fluorescent
lamps for example, at distances corresponding to one half of a step, an ellipse
is produced. An example of such an ellipse is shown on a large scale in
Fig. 6.17. The isotemperature lines touching the ellipse pass through the
points 4 140 K and 4 480 K.

Fig. 6.17. Tolerance ellipse for 'White'


lamps.
BBL = black body locus.
W = chromaticity of 'White' lamp;
colour temperature 4300K.
;-----~~;--------;0.;-;!:B A = isotemperature line for 4140K ..
- x B = isotemperature line for 4480 K.

As long as the chromaticities of any two lamps of the colour 'White' fall
within this ellipse, the separation is less than one step and the colour dif-
ference will be acceptable in practice. Statistical methods have been used to
show that with modern production methods the manufacturing tolerances
can be made small enough to ensure that practically all lamps fall within
the one step ellipse. It is even possible to be somewhat more strict for de-
viations in the direction perpendicular to the black body locus. Similar
ellipses can be plotted for the colours 'Daylight' and 'Warm White' and for
these also the colour points of the lamps can be made to fall within the one
step tolerance.
These considerations on colour tolerances and the data given in the pre-
ceding paragraph on colours of fluorescent lamps all concern 40 W lamps
of 37 mm (1-t in) bulb diameter, operating on a current of 430 rnA at 25
oc (80 F) ambient temperature. Lamps operated at a higher current den-
sity or with a higher ambient temperature show a slight colour shift to-
wards the green. This shift is due to the fact that, with higher bulb tern,
peratures and consequently higher mercury vapour pressures, the intensity
of the visible spectrum of the mercury discharge, and notably that of the
COLOUR AND COLOUR RENDERING 91
green line ( = 546 nm), increases in comparison with the light from the
phosphor.
During the life of a fluorescent lamp there is no appreciable change in
colour.

6.8 Colour rendering


So far we have only dealt with the colour appearance of fluorescent lamps.
The function of a lamp, however, is to illuminate the objects around rather
than to be seen itself. The colour of an object is to a large extent dependent
on the characteristics, and especially on the spectral energy distribution, of
the light in which it is viewed. In order to understand why this should be so,
let us first study the factors to which the colours of objects are due. An object
that reflects or in fact disperses all the light falling on it, gives the impression
of being white. Such objects do not remove any radiation from the visible
spectrum and it is said, by definition, that the reflection factor (! at all wave-
lengths is 1. Thus the colour appearance of a white object and that of the
lamp are the same. A thick layer of magnesium oxide has this property to
a high approximation and thus looks pure white. In Fig. 6.18 reflection
factors are plotted as functions of wavelength.

/
0.8 /
I
I
e I
I

/2
0.$1- - - + - - - ~,.~--~
/ 1\
I I '.
3/ :
j
'\
I
-.., . I \
\
4 // \ ./ /
.....---- .
,. . . \ I
/ \
\

0..2
/ \\ / \

I Y-.. ._
/ / . " ...... . ...... -r
1.2.3 _,/
/
/
I
............... ~-~:~:.::::7 /
~oo=--------~~~-----L----~6~00~----------~ro~o----~
). r~m)

Fig. 6.18. Reflection factors of white and coloured objects.


A. = wavelength in nm.
11 = reflection factor.
11 = 10 white object as approximated by magnesium oxide layer. Curve 1 = red object.
2 = yellow object. 3 = green object. 4 = blue object.
The measurements reproduced were made on printing inks, but for the sake of clarity
the curves have been somewhat simplified.
92 FLUORESCENT LAMPS
I.O violet blue green yellow orange red
p T I I I I

-
0.8

v
0.6 ~

0.4 I

' ~"----[7I
f---..-
0.2 Fig. 6.19a. Reflection curve
of a purple object.
e = reflection factor;
0 A = wavelength in nm.
400 500 fXlO 700
-JrnmJ
E ! I

I ,-- /
100
_,-"1

8 ;
;
"'"'
' ;
;
;

60 / --
!.---- r-_ ,"'
/

40
I/ ;
; r-- 1---
~
, "' Fig. 6.19b. Spectral energy distribu-
/' tion curves.
20
,,"
1-----------
E = relative radiant power emitted at
1--",. wavelength A. Full line for daylight.

_li
0 Dotted line for incandescent lamp.
400 500

160

140

120 --- ----


/
/
/
~/
100 -~ 1'---
/,
/I I
I
80 f-- -
I
I
I

60 1----------- 1-
:v- -
- -- -----

I Fig. 6.19c. Spectral energy distribu-


tion curve for the light reflected by

U=l~+=--
40 --
the object of Fig. 6.19a.
lr E = relative radiant power reflected

-~
at wavelength A.
2 or---- . Full line in daylight illumination .
-............ 1:::_ __
,_.. ....
'
I
I
Dotted line in incandescent lamp-
light illumination.
500
-!!!E. .< (nm) 700
COLOUR AND COLOUR RENDERING 93
In the case of coloured objects the reflection factor is not independent on
wavelength, as can be seen from the examples given in Fig. 6.18. Objects
reflecting red, orange and yellow light, but removing or absorbing green, blue
and violet, are red in colour; if red, orange and yellow are reflected, and green
as well, then the colour is yellow. Green objects absorb red, orange, blue and
violet rays and reflect mainly green light, whilst blue objects absorb yellow,
orange and red. The increase in reflection in the very deep red, such as occurs
in the two last-mentioned colours of Fig. 6.18 at wavelengths above A.=
700 nm, has no appreciable effect on the colour, since, in accordance with
Table 6.1, all three photo-electric cells representing the eye are almost
insensitive to these rays. Purple is not a simple colour; the reflection curve
reveals two maxima, one in the red region and another in the blue. An
example of such a curve is reproduced in Fig. 6.19a.
The light striking the eye after reflection from a given coloured object,
say a purple one, will have different spectral characteristics when the object
is illuminated successively by two light sources of different colours. Extreme
instances of this are incandescent lamp light and daylight. With other, less
diverging types of light, the results are similar but not so pronounced. If a
purple object (Fig. 6.19a) is illuminated first with daylight of which the
spectral composition is as given by the full line in Fig. 6.19b, the spectral
composition of the reflected light will be found by multiplying at each wave-
length the energy falling upon the object (Fig. 6.19b, full line) by the reflec-
tion factor taken from Fig. 6.19a. The result is as shown by the full line in
Fig. 6.19c. If the process is then repeated for incandescent lamp light, the
spectral energy distribution of which is given in the dotted line of Fig. 6.19b,
the result will be as depicted in the dotted line of Fig. 6.19c, which is quite
different from the full curve. The difference between the two spectral distri-
butions means that the chromaticities of the purple object are widely different
for the two kinds of light. Calculation shows that with daylight illumination
x = 0427, y = 0259 whereas for the incandescent lamp x = 0570, y =
0331. We might introduce a measure for this colour variation by assessing
the number of colour steps it corresponds with, but this would be pointless
in view of the fact that the tints do not even resemble each other. Thus the
enormous difference between the spectral energy distributions of daylight
an.d incandescent lamp light has the consequence that the two types of light
cannot be used together without producing strongly coloured shadows or
fringes. This conclusion is the same as that inferred above from the colour
difference between the light sources themselves.
It is important to note here that, in spite of the large differences in chro-
maticities we found, the colour perceptions of most objects around us in incan-
descent lamp light are substantially the same as in daylight, provided that
incandescent lamps are the only light sources present in the room where the
observations are made. This phenomenon, called the persistence of colours,
is a consequence of chromatic adaptation of the eye. There is a relatively
large group of light sources, differing in spectral energy distributions and
consequently giving different colour points for a given coloured object,
in which, owing to chromatic adaptation, almost every pigment very nearly
gives the impression of its daylight colour, so that these light sources can
be said to have 'natural' colour-rendering properties. To this group belong
94 FLUORESCENT LAMPS

- as every day experience tells us - daylight in all its varieties (Tc =


5 000-20 000 K) and incandescent lamp light (Tc = 2 400-3 000 K).
Experiments have shown that the results of chromatic adaptation can be
fairly accurately predicted. The state of adaptation appears to be determined
mainly by the colour of the light source *. Application of the rules deduced
from the experiments showed that the radiations of the black body in the
whole range from 2 400 K-20 000 K have natural colour rendering pro-
perties.
In the preceding sections we did not consider the fact that light sources
possessing entirely different spectral energy distributions may have the same
colour point. As the state of chromatic adaptation will be the same in both
cases the perceived object colours must then necessarily become different.
An example is illustrated in Fig. 6.20.

TOO

1~ b

0 400 500 600 700


- - - A (nm)
Fig. 6.20. The light source having the energy distribution (a) the so-called equal-energy
spectrum shows the same white colour as that with the distribution (b). Both light sources
also have the same luminous flux and the same colour point x= t, y= t; i.e. the 'white
point' in the mixture diagram.
E = relative radiant power emitted at wavelength A.

The radiation a, the so-called equi-energy spectrum has the colour point
x = t, y = t, and makes the impression to be white. Radiation b, con-
taining only the two wavelengths A = 589 nm and A = 486 nm, the latter
with twice as much energy as the former, has the same colour point and,
therefore, gives the same white impression. Object colours generally are
rendered entirely differently by the two kinds of light. In radiation a all objects
show their natural colours. In radiation b, the rather light red pigment having
the spectral reflection factor given in curve 1 of Fig. 6.18, looks yellow. If
the red is deeper, the colour in radiation b is brown. In radiation b only
yellow and blue pigments show approximately their natural colours. The
* SeeP. J. BOUMA and A. A. KRUITHOF, Philips Techn. Rev. 9, 2 (1947) and Philips Techn.
Rev., 9, 257 (1947).
COLOUR AND COLOUR RENDERING 95
colour rendering properties of radiation b are therefore very poor, those of
a good or natural.
We now return to fluorescent lamps. It follows from the example given
above that, although the colour points for the lamp colours have been fixed,
the spectral energy distributions are not all determined in consequence.
Therefore phosphors having widely different spectral energy distributions
can be used as components of the mixtures. This degree of freedom in prin-
ciple may be employed to achieve two more or less opposite aims.
(1) The highest possible light output.
(2) The best possible colour rendering properties.
For the 'standard' fluorescent lamps dealt with in Section 6.6 the former
alternative was given preference. Halophosphate phosphors are particularly
well suited for these lamps since the long wave fluorescence is mainly con-
centrated in the yellow spectral region, where the eye is most sensitive.
For the 'de Luxe' lamps, however, the emphasis has been placed upon the
other possibility, i.e. good colour rendering. As all kinds of intermediate
cases can be realised by varying the composition of the phosphor mixture
it will be clear that it is highly desirable to have a method that allows us
to evaluate the quality of the colour rendering. In recent years a large amount
of work has been done in order to establish such a method. It will be dis-
cussed in the next section.

6.9 The specification of the colour rendering properties


In order to relate a difference in spectral energy distribution to colour ren-
dering various methods have been proposed, all based on two principles that
were discussed by Bouma as early as 1939.
In the first approach, the spectral band method, the spectrum is divided
into a number of bands, and the contributions in these bands are compared.
These contributions may be expressed directly in terms of energy, or the
energy may be multiplied with a weighting function. In the method recom-
mended provisionally by the C.I.E. in 1948, which was identical with the
one proposed by Bouma in 1939, the visibility curve V(.A) was used as a
weighting function, so that the amounts of light of the bands were com-
pared. This band system consisted of three narrow bands in the blue and
five broader bands in the rest of the spectrum. It appeared that the use of
the 1948 spectral band method for the quantitative evaluation of light sources
presented various difficulties that could only partly be remedied by choosing
another band division and a more satisfactory weighting function, for in-
stance one having three maxima.
The second approach resulted in a method which is now officially recom-
mended by the C.I.E. (Publication No. 13 (1965)). It is based on an entirely
different principle; the behaviour of a number of test colours. It was shown
that the average of the colour differences that occur when a number of test
colours are illuminated successively by the lamp to be tested and by a refer-
ence illuminant with natural colour rendering provides a satisfactory measure
for the colour rendering properties of the light source. This average colour
96 FLUORESCENT LAMPS

difference appeared to be only slightly dependent on the choice of the test


colours as long as these are of comparable saturation and evenly distributed
over the main hues: red, yellow, green, blue and purple. With strongly
saturated test colours the colour differences are greater than for less saturated
ones, but there exists a high correlation between lamp ratings based on
different sets of test colours.
In the C.I.E. Test Colour Method 1965 eight test colours of medium satura-
tion are used. The chromaticity of a test colour is calculated from the spectral
energy distribution curve of the light source to be tested and the spectral
reflectance curve of the colour. The same is done for the reference illuminant,
and the colour difference is calculated. This cannot easily be done directly
from the C.I.E. x, y diagram, since in this diagram equal distances do not
generally correspond to equal colour differences. Therefore the x, y co-ordi-
nates are transformed into the co-ordinates u, v of a so-called uniform-
chromaticity-scale diagram for which this correspondence applies reasonably
well. There are a few of these diagrams; the u, v diagram proposed by
MacAdam (1937) was adopted by the C.I.E. These co-ordinates are related
to x and y in the following way
4x 6y
u=------------ v = -----------
-2x + 12y + 3 -2x+12y+3
The colour difference is obtained from the distance between the chroma-
ticities in this diagram or calculated from L1E1 = 800 (V(L1u 1) 2 + (L1v;)2 in
which L1E1 is the colour difference for the test colour considered expressed
in NBS units of colour difference, and .du1 and L1v 1 stand for the differences
in the u and the v co-ordinates respectively. L1E1 is then used to calculate
the special colour rendering index by means of the formula
R 1 = 100- 46 L1E~o
Finally, the General Colour Rendering Index Ra is obtained as the average
of the eight R 1 values. The scale has been so chosen that the reference illu-
minant was assigned a General Colour Rendering Index 100 (All L1E1s are
zero) while a warm white fluorescent lamp studied in the course of the investi-
gations was rated at 50. This scale is therefore largely arbitrary, its physical
basis follows from the relation R 1 = 100- 46 L1E1, which shows that 46
points in the scale correspond to one NBS unit of colour difference. The
NBS unit of colour difference is noticeable only in case of a direct com-
parison, therefore R 1s exceeding 95 indicate that in practice the colour ren-
dering of the relevant light source cannot be distinguished from that of the
reference illuminant.
In its present form the General Colour Rendering Index must primarily
be considered as a means to check a light source against a reference illuminant
the colour rendering properties of which are accepted to be natural by de-
finition. For daylight lamps natural daylight is the obvious reference illu-
minant. The spectral composition of natural daylight varies considerably,
however, the colour temperature ranging from 5 000 K-50 000 K. As the
reference source should have about the same colour temperature as the lamp
to be tested, a series of natural daylights (the so-called Reconstituted Day-
COLOUR AND COLOUR RENDERING 97
lights) has been standardised for this purpose. They are based on measure-
ments in various parts of the world.
The choice of the black body radiator as a reference illuminant for light
sources of 5 000 K and lower is less obvious. Although chromatic adap-
tation causes the colour perceptions of object colours to be closely similar
under tungsten light and in natural daylight, they are not exactly the same.
Nevertheless, whether by nature or by habit, we accept tungsten light as a
satisfactory light source, whereas a warm white fluorescent lamp with Ra
deviating markedly from 100 definitely conveys the feeling that many objects
are shown in wrong colours. This justifies the choice of a black body radiator
as a reference source for the lower colour temperature region. On the other
hand it must always be kept in mind that the Ra data are based on such a
low temperature black body radiator as a reference. Therefore the Ra is not
an absolute entity; if a daylight lamp and a warm white lamp both have an
Ra of 98, this does not imply that their colour rendering properties are
identical. As both values are close to 100 the colour rendering will be dif-
ferent to the same extent as those of tungsten light and natural daylight.
The General Colour Rendering Index should not be considered as a magic
number that describes the colour rendering properties of a light source un-
ambiguously. It is based on the average of a number of colour differences,
and as such it does not carry information on the behaviour of the individual
test colours. Only if the average of the colour differences is small, such as for
lamps with Ra > 90, it becomes very unlikely that any of the colours should
show serious colour distortion. But as Ra becomes smaller its significance
decreases correspondingly. All that can be said about a lamp with Ra = 75
is that there will probably be many colours that will show noticeable colour
shifts.
But despite this restriction the General Colour Rendering Index is a very
useful tool in selecting the most suitable lamp for a given application. For
critical colour inspection purposes, such as colour matching, lamps must be
chosen with an Ra value of at least 90, preferably 95 or even higher. For office
lighting, shop lighting, etc., values of 80-85 will generally prove satisfactory.
The significance of differences in the index Ra must not be overestimated.
Lamps having the same general kind of spectral energy distributions should
have a difference of at least 2 units before the difference may become notice-
able in lighting practice. Also, the fact that two lamps of about the same
colour have the same Ra value does not necessarily mean that their colour
rendering properties are identical. For instance, one may have an excess of
deep red rediation over the reference radiator and the other may show a
deficit in the same region. As the spectral energy distribution will generally
be published together with the Ra value, cases like this will be easily re-
cognised.
98 FLUORESCENT LAMPS

6.10 The shortcomings of the 'standard' fluorescent lamps


In trying to improve the colour rendering properties of fluorescent lamps
the spectral energy distributions of the standard types as given in Figs 6.13,
6.14 and 6.15 may be taken as the starting point. The most important devia-
tions from either natural daylight and or from the corresponding black body
radiators which, as we have seen in the preceding section, are taken as refer-
ence sources, are similar for all three lamp colours:
(I) A deficit in the extreme-red spectral region.
(2) Two excesses of power over the reference source, namely in the blue-
violet (the mercury line A. = 436 nm) and in the yellow-green (between
A. = 540 nm and A. = 590 nm). The two excesses are interdependent. In
order to compensate for the unavoidable blue-violet mercury line the
mixture has to be chosen such that an excess in the complementary colour
of A. = 436 nm is obtained. Strictly speaking the complementary colour
for the mercury line is different for the three types of fluorescent lamps.
It can be determined by drawing a straight line through the colour point
of the mercury line and the colour point of the lamps to the opposite
side of the spectrum locus. The intersection point of this line and the
locus is the complementary colour for A. = 436 nm. For all three lamp
colours one arrives then in the yellow-green spectral region.
It is easy to understand in general how the two deviations affect the colour
rendering qualities of the lamps. The deficit in the red causes red objects to
be rendered poorly and vivid red colours acquire a dull aspect. The devia-
tion mentioned under 2, although giving blue objects a definite violet tinge
is especially detrimental for the rendering of yellow objects. The reason is
that yellow objects absorb the blue-violet excess compensating the yellow-
green excess in the spectrum of the light source and also for objects reflecting
both radiations, such as for example a white table cloth. The result is the
well-known green discoloration shown especially by poorly saturated yellow
colours when lighted by 'standard' fluorescent lamps. In this respect the
behaviour of many foodstuffs, such as butter and potatoes, is notorious.
They often acquire a slightly unappetising appearance with this illumination.
The deficit of red radiation is of equal importance for all three lamp
colours. The effect of the blue-violet mercury line is more serious in the case
of 'Warm White' and 'White' lamps than with 'Daylight' lamps. This is
also easily understood: the contribution of the mercury light and therefore
also that of the blue-violet mercury line to the total luminous flux is about
the same for all lamp colours. The light of the 'Warm White' lamp contains
only a small amount of blue radiation which consequently is largely concen-
trated in the blue-violet line; the light of the 'Daylight' lamps contains con-
siderably more blue radiations so that the mercury line does not play such
an important role. The position of the 'White' lamp is intermediate between
the two other types.
In the following section we shall consider the possibilities to eliminate the
two shortcomings mentioned.
COLOUR AND COLOUR RENDERING 99
6.11 The development of de Luxe lamps
Starting from the 'standard' types of fluorescent lamps made with halo-
phosphates the deficit in the red may be remedied by the addition of a phos-
phor giving deep red radiation. Magnesium arsenate or magnesium germa-
nate may be used for this purpose, both having narrow emission bands with a
maximum at 656 nm. If these phosphors are added to a halophosphate, or to
a mixture of a white and a blue halophosphate, the colour point will be
displaced in the direction of that of magnesium arsenate or germanate,
that is to say towards higher values of x and lower values of y. In order to
return to the correct colour point it will be necessary to add a blue-green
phosphor or a phosphor mixture, such as a blue halophosphate and green
willemite. As an alternative, instead of using a halophosphate, another long
wavelength phosphor such as a tin-activated orthophosphate may be used
as a starting point and a suitable blue phosphor, or a mixture of a blue and
a green phosphor may be added to get the desired colour point. It is difficult
to predict which will be the most attractive way to reach our goal: good
colour rendering properties without losing too much in luminous efficacy.
Increasing the amount of deep red radiation automatically involves a drop
in lumen output.
Even with a restricted number of available phosphors a systematic experi-
mental investigation of the various possibilities amounts to a formidable
task; a complicating factor being that the answer to the question of which
lamp is to be preferred with regard to the colour rendering properties is a
matter of the personal preferences of the observers. Another approach which
has proved to be very useful was followed by the authors *. It was stated
in Section 6.5 that with three given components of the phosphor mixture
the proportions of their contributions to the light can be calculated, as soon
as the chromaticity of the lamp is given. With four components the ratio
between the contributions of two of them, for instance of the two long wave-
length phosphors, may be chosen beforehand and with that fixed ratio we
can then proceed as in the case of the three component mixture. Once the
contributions of the components are known the spectral energy distribution
of the lamp can be calculated. The final step is to calculate the colour shift
of a number of test colours. In case one of the phosphors absorbs visible
radiation, such as magnesium arsenate which absorbs in the deep-blue, a
correction can be applied to the spectral energy distribution curves of the
other components.
By means of an electronic computer this kind of computation is easy to
perform and since there is now available a standardised method for evalua-
ting the colour rendering properties resulting from the various spectral energy
distributions, this computational approach is a very attractive one, which has
already proved its usefulness. Without entering into details the results can
be briefly summarised as follows.
As was pointed out already in Section 6.10 the effect of the blue-violet
mercury line is easiest to compensate for in the case of the daylight lamp.

* A. A. KRUITHOF and J. L. OuwELTJES, Philips Techn. Rev. 18, 249 (1957).


100 FLUORESCENT LAMPS

There is a sufficient choice of blue emitting phosphors such as calcium halo-


phosphate, strontium halophosphate, etc., to find a satisfactory solution, and
consequently most of the lamp makers have succeeded in developing day-
light lamps with a colour rendering index of 90-95. The most interesting
case is provided by the cool white lamp, with a colour temperature of
4 000-4 500 K. With phosphors such as the blue calcium halophosphate or
strontium halophosphate it is not possible to make lamps with Ra > 90,
but by absorbing part of the blue-violet mercury lines lamps have been de-
veloped with Ra = 95 or even higher (see Section 6.12). This absorption is
effected by means of a separate coating of magnesium arsenate applied
directly to the glass. The 254 nm radiation is absorbed by the inner coating
of phosphors, whereas the long wave ultra-violet, the blue-violet mercury
lines and the continuous radiation from the phosphors in this region are
attenuated by the magnesium arsenate layer. If the quantity of magnesium
arsenate necessary to absorb the blue-violet to the required degree would
be mixed with the other phosphors its emission would be much stronger
than desirable.
It is interesting to note, however, that the excess of radiant power due to
the mercury lines can in principle satisfactorily be compensated for by using
a blue phosphor the emission band of which is narrower than that of the
calcium or strontium halophosphates. As an example barium strontium di-
silicate activated with divalent europium may be mentioned. With this and
similar phosphors it was found to be possible to design cool white lamps
with Ra ~ 95, without applying an absorbing coating.
For the warm white lamps, having a correlated colour temperature of
about 3 000 K, there is no hope of ever finding a blue phosphor that will
make it possible to design a lamp with excellent colour rendering properties
without having recourse to a blue absorbing layer. But by applying a magne-
sium arsenate coating (the only substance used for this purpose so far) and
selecting suitable phosphors a lamp was developed (see Section 6.12) with
an Ra value of 93. As might be expected the light of such a lamp can
hardly be distinguished from incandescent lamp light, and consequently it
is rapidly finding application in home lighting, restaurant lighting, etc.

6.12 Data for some of the Philips de Luxe lamps


6.12.1 Colour 57. Daylight 7 400 K. x = 0301, y = 0311
General colour rendering index 93
For colour inspection purposes such as occur in the textile industry, lamps
with a high colour temperature are generally preferred. Another point to be
considered is the frequent use of fluorescent dyes (for instance the so-called
optical bleach), which makes it desirable to increase the ultra-violet output
of the lamp. In order to match natural daylight also in this respect an ultra-
violet emitting phosphor has been added to the phosphor mixture. The spec-
tral energy distribution is shown in Fig. 6.21.
COLOUR AND COLOUR RENDERING 101

E/IP

i 200

100
Artificial
daylight lamp

300 400 500 500 700


- - - "A(nm)
Fig. 6.21. Spectral energy distribution curves for Philips artificial daylight fluorescent
lamp (Colour 57) and reconstituted daylight 7 500 K.
A = wavelength in nm.
E/ rp = energy flux in I 0- 6 W for a wavelength band of I 0 nm width per lumen light flux.

6.12.2 Colour 55. Daylight 6 500 K. x = 0316, y = 0334


General colour rendering index 93
In other cases there is no specific interest in an increased amount of
ultra-violet or even the desire to keep the ultra-violet as low as possible be-
cause of radiation damage, the colour of the lamp and the colour rendering
properties still matching those of natural daylight. This application is covered
by colour 55, the spectral energy distribution of which is closely similar of
that of colour 57 except in the ultra-violet.

6.12.3 Colour 34. Cool White de Luxe. x = 0384, y = 0368


General colour rendering index 86. Colour temperature 3 800 K
Daylight lamps are not suitable for lighting offices, shops, etc. unless
there is a sufficient access of natural daylight, so that a high illumination
level is maintained. As this is a requirement that will be complied with only
in exceptional cases, a lamp with a lower colour temperature will in general
offer a better solution. Although for this kind of application no severe requi-
rements are made with regard to colour rendering, the colour distortions as
they occur with the Standard Cool White fluorescent lamp (especially that
of the human skin) have proved to be objectionable, and have led to the
development of Cool White de Luxe lamps which represent a fair com-
promise between good colour rendering properties and high luminous effi-
ciency. Colour 34 belongs to this category of lamps. Its spectral energy dis-
tribution is given in Fig. 6.22.
102 FLUORESCENT LAMPS

!~

I
200
Full
-raCI7a1iir

100

--
400 500 600 700
-f..(nm)
Fig. 6.22. Spectral energy distribution curves for Philips White de Luxe fluorescent lamp
(Colour 34) and full radiator 3 900 K.
A = wavelength in nm.
E/rp = energy flux in 10- 6 W for a wavelength band of 10 nm width per lumen light flux.

6.12.4 Colour 27. Softone. x = 0464, y = 0409.


General colour rendering index 94. Colour temperature 2 700 K
For a long time there has been a general need for a fluorescent lamp that
would be suited for the lighting of homes, restaurants and the like, where
the desire for intimacy implies a lower illumination level and a low colour
temperature. As incandescent lamps are very often used for additional local
illumination such a fluorescent lamp should not only have a colour appear-
ance but also colour rendering properties that are closely similar to those of
incandescent lamps. Both requirements are fulfilled by the recently developed

!~
,-"'
,/Full
--
/ radiator

t 200

100

400 500 600 700


-f..(nm)

Fig. 6.23. Spectral energy distribution curves for Philips 'Confort de Luxe' fluorescent
lamp (Colour 27) and full radiator 2 700 K
A = wavelength in nm.
E/rp =energy flux in 10- 6 W for a wavelength band of 10 nm width per lumen light flux.
COLOUR AND COLOUR RENDERING 103
Colour 27, the spectral energy distribution of which is shown in Fig. 6.23.
In accordance with the analysis given in the preceding section it was found
necessary to suppress the blue mercury lines and the lamp has a double
coating.

6.12.5. Special lamps.


In addition to the lamps just mentioned a few others have been devel-
oped for special purposes. The Graphic Arts Industry has adopted a light
source of 5 000 K for inspection of colour prints, colour transparencies,
etc., a general colour rendering index of about 95 being desirable (Philips
Colour 47). For hospital and museum lighting a lower colour temperature
is preferred, the requirements with regard to colour rendering being the
same. Colour 37, (Trucolour) with a colour temperature of 4 200 Kanda
general colour rendering index of 96 covers both requirements. Colour 32,
with a colour temperature of 3 000 K, and a general colour rendering index
of 85 may be considered as a compromise between high luminous efficiency
and colour rendering in the warm white region.
This survey of existing lamp types is given to illustrate the various types
available at the moment rather than to describe lamps that are to remain
unaltered in the future. On the contrary, it is expected that new phosphors
will open the way to further improvements, especially to an increased light
output for those lamps that have already excellent colour rendering pro-
perties.
Chapter 7

Stabilisation of the discharge


Th. Hehenkamp

7.1 Introduction
It has already been shown in Chapter 4 that gas-discharge lamps have a
negative voltage-current characteristic; as a consequence the lamp current
is not limited if the lamp is connected directly to a constant voltage supply
source. It is therefore necessary to connect an impedance in series with the
lamp, in order to restrict the current to the desired value.
A second consequence of the negative voltage-current characteristic ap-
pears only when alternating current supply is employed. If, therefore, the
lamp is fed with a current of sinusoidal wave form, the lamp voltage is
considerably distorted. Fig. 7.1 shows the form of the lamp voltage during
one half cycle, as constructed from the current v. voltage characteristic of
the lamp. If the lamp is symmetrically constructed, the same voltage pattern
will appear in the following half cycle but with opposite sign.
For frequencies higher than a few tens of cycles per second, the construc-
tion shown in Fig. 7.1 is no longer valid, since the given voltage-current
characteristic only applies to slow current variations (static characteristic).
With more rapid current changes the characteristic alters, because the ionisa-
tion conditions in the gas discharge can no longer follow the variations in
the current. This effect leads, at increasing frequencies, to a lamp voltage
STABILISATION OF THE DISCHARGE 105
Vz

lz
I I
I
\ I I
I I \
I

' ' '-...."-


I I \
I I I I

\ '
..... _
I
\ I
\ I
\ \
\
'
' " "' ' '
---
'''
' -- ---
Fig. 7.1. Diagram showing lamp voltage during one half cycle of sinusoidal alternating
current, as constructed from the current v. voltage characteristic of the lamp.

pattern which more and more closely approximates to a sinusoidal shape,


provided that the current wave-form supplied is also sinusoidal. The mini-
mum frequency required to achieve a sinusoidal voltage pattern depends on
the lamp diameter, length, current and filling-gas pressure. A smaller dia-
meter, smaller length, a higher current and lower gas pressure demands a
correspondingly higher frequency to achieve the same effect. This behaviour
is illustrated by the oscillograms shown in Fig. 7.2, which have been taken
from different lamps and at various frequencies.
If, in a given case, the frequency is sufficiently high to result in a sinusoidal
voltage for the lamp under consideration, it is possible to calculate the re-
quired ballast characteristics by the use of standard electrical formulae and
vector diagrams. At the more usual frequencies of 40-60 Hz, however, the
influence of harmonics in the voltage and current is so great that the use
of vector diagrams demands considerable caution and results only in a rough
approximation to the correct ballast requirements.
More accurate calculation is often facilitated at these frequencies by repre-
senting the lamp voltage schematically by a square wave* (Fig. 7.3), as
will be used in a few cases in Section 7.3.

t According to Fourier, the square wave can be developed in the series:


v
f(wt) = 4 _!_(sin wt + t sin 3wt + gsin 5wt . ...... etc.) (7.1)
n

* E. G. DoRGELO, 'A.C. Circuits for Gas Discharge Lamps', Philips Techn. Rev.'2, 103-109
(1937).
106 FLUORESCENT LAMPS
'TL' 40W 'TL' 20 W 'TL'D 30W
T 12 - 48" - 044 A T 12 - 24" - 039 A T8-36" - 037 A

50 Hz

100Hz

200Hz

800Hz

8000Hz

Fig. 7.2. Oscillograms of lamp voltage of 6 different types of lamp at frequencies of


50, 100, 200, 800 and 8 000 Hz.

so that the lamp can be considered in this case as a generator which supplies
an opposing voltage to the mains voltage containing all uneven harmonics
of the mains frequency. Since the actual lamp voltage pattern shows no
STABILISATION OF THE DISCHARGE 107
'TL'D 15 W 'TL' 13 W 'TL' 4 w
T 8 -18" - 032 A T 5 - 21" - 016 A T 5 - 6" - 015 A

~. ~ ~.~~~~
... ....

.~ ~ 50 Hz
- . . . .

100Hz

200Hz

800Hz

J\/\) /\/\/ 8000Hz

sharp corners, the series can be disregarded after the first few terms.
If the lamp is somewhat asymmetric, due to a difference in the emissivity
of the two electrodes, then the even harmonics can also appear, only the
second harmonic being of any significance, however. Further reference to
the consequences of the even harmonics will be made in Section 7.4. t
108 FLUORESCENT LAMPS

Fig. 7.3. Lamp voltage, simplified to a square wave. The square wave can be deve-
loped into a series of sine terms. The first three terms of this series and their sum are
drawn in the diagram.

Depending on the nature of the impedance used as a stabilising ballast,


the deformation of the lamp voltage will result in a certain deformation of
the lamp current. Moreover, as a consequence of the deformation of the
lamp voltage, the actual power taken up by the lamp will be less than the
apparent power, so that the power factor of the lamp is less than unity and
can be represented by the expression:

(7.2)
II VI
If v1 is constant during a half cycle and the current can be assumed to be
sinusoidal, the above equation can be written:

V2
---;; I 1 V1 J"sin wt dwt
0
IX1 = - - - - - - - - - = 09
II VI
This value of the power factor is usually approached fairly closely in practice.
Since the divergence from unity is not due to a phase difference between the
lamp current and lamp voltage but to the distortion of the lamp voltage,
it is more accurate to speak of distortion factor or form factor than of
power factor.
STABILISATION OF THE DISCHARGE 109
If we disregard more complicated circuits and confine ourselves in this
chapter to single circuit elements in series with the lamp, the following pos-
sibilities arise:
I. Resistor, with little or no temperature coefficient of resistance.
2. Resistor with high temperature coefficient of resistance.
3. Self-inductance.
4. Capacitor.
The first two elements can be used for both d.c. and a.c. supplies, the
other two only on a.c. supplies.

7.2 Direct current supply


7.2.1 Resistors with small temperature coefficient of resistance
The lamp voltage and current required for a given power in the lamp can
be determined from the lamp characteristic. The difference between lamp
voltage and the mains voltage, or, in the general case, the no-load voltage
V0 , must be taken up by the resistor. The power dissipated thereby in the
resistor is useless for the light production, so that the efficiency is consider-
ably reduced. In order to keep the efficiency as high as possible the no-load
voltage should be chosen as close as possible to the lamp voltage; this
reduces the stabilising effect of the resistor and thus increases the influence
of fluctuations in mains voltage. The variation of lamp wattage (i.e. for
practical purposes the variation in light output) is shown in Fig. 7.4 as a
function of the ratio of lamp voltage V1 to the no-load voltage V0 , for a
'TL' 40 W lamp.
Another consequence of a small difference between no-load voltage and
lamp voltage is that the current flowing through the resistor under short
circuited lamp conditions becomes very high. This short circuit current I. is

4 klz at ~TO%
100
%
80 I

I

60

. "'~~

v
40 /

Fig. 7.4. Percentage variation in lamp


power as a function of V,/ V0 on voltages
of 90% and 110% of the nominal value,
20
- ~

for the case of a lamp operated on 0


direct current with a resistor as series- 0.4 0.5 0.6 0.7 0.8 0.9
impedance. -Vzflil
110 FLUORESCENT LAMPS

Isjzz
5.----.----.----.-----.--~

Fig. 7.5. Efficiency and l,/11 curves as


~--,o~,a.--~a~4~--~~~6--~~~B--~t2 a function of VtfV0 , for the case of a
lamp operated on direct current, with
-11/Vo a resistor as series-impedance.

important if a starter is used to strike the lamp which short circuits inter-
mittently during striking in order to pre-heat the electrodes (Fig. 7.5). For
these two reasons the no-load voltage can normally not be made less than
twice the lamp voltage, so that the efficiency of the circuit is reduced to 50%.

t The value of the resistance R is determined by:


V0 - Vt
R=--- (7.3)
It
The short circuit current I. = V0 /R, so that, with the help of equation 7.3
there follows:
I. 1
(7.4)
I, v,
1--
Vo
The efficiency 'YJ of the circuit is equal to the lamp power divided Hy the total
power, hence:
V1 I 1 V1
'Yj=-=- (7.5)t
Voii V0

7.2.2 Resistors with high positive temperature coefficient of resistance


A considerable improvement in efficiency can be achieved by using a
resistor of high positive temperature coefficient as ballast. The sensitivity to
mains voltage variations is then much smaller, for a certain ratio of lamp
STABILISATION OF THE DISCHARGE Ill
voltage to mains voltage, since the change in resistance counteracts the cur-
rent variations caused by the fluctuations in the mains voltage. The largest
effect is obtained with a current stabilising valve, e.g. iron-hydrogen resis-
tor *, which, with a correct choice of the regulation range, can keep the
lamp current practically constant, even with large variations in mains voltage.
To keep the lamp current constant with mains voltage variations ranging
from 90% to 110% of the nominal value, the regulator valve must operate
over a range G (i.e. the ratio of the highest to the lowest voltage over the
valve, at constant current) of:
11 V0 - V1
G=----
09 V0 - V1
Values up toG= 6 are possible in practice, so that a ratio VtfV0 = 086
is attainable. During the ignition of a fluorescent lamp, the full mains voltage
is applied momentarily across the regulating valve. To prevent damage to
the latter it is desirable that the current does not exceed the regulation range
too much; the value of VtfV0 , thus of r;, is therefore limited in practice to
07-075.
Disadvantages of the iron-hydrogen resistor are: too small a resistance to
shock and the dangerous nature of the hydrogen gas filling. A more practical
solution, with a more robust construction and almost equally favourable
properties, is obtained by locating a tungsten spiral in a mixture of hydrogen
and nitrogen**. An example of what can be achieved with this combination
is shown in Fig. 7.6.

.d 11.1 at l1n 10%


TOO
O/o
t80
0.9Vm
60 1/
1/
I

~
Fig. 7.6. Percentage variation in lamp Vl.TVm
power as a function of VtfV0 on volt-

-
ages of 90% and 110% of the nominal 20 0.9Vm
value, for the case of a lamp operated
on direct current with a stabilisation
~ v,,'
;;_ ....... T.lVm

tube (dotted line) and with an incan- =F==-== -:.::.=


0
descent lamp (full line) as series-impe- G.4 0.5 0.6 0-7 0.8
dance. -\'1/Vo

* J. G. W. MULDER, 'Current-regulator tubes', Philips Techn. Rev. 3, 74-80 (1938).


** L. P.M. TEN DAM and D. KoLKMAN, 'Lighting in Trains and other Transport Vehicles
with Fluorescent Lamps', Philips Techn. Rev. 18, 11-18 (1956/57).
112 FLUORESCENT LAMPS

Another temperature dependent resistor which can be used as stabilising


element is an incandescent lamp. To avoid overloading the incandescent
lamp filament during the ignition of the fluorescent lamp it is desirable to
select an incandescent lamp suitable for use on the mains voltage upon which
the fluorescent lamp must operate; the incandescent lamp is then consider-
ably underloaded under operating conditions.
Results of measurements made on a 'TL' 40 W lamp with an incandescent
lamp in series are included, for comparison purposes, in Fig. 7.6. The favour-
able influence of resistors with high positive temperature coefficient of resis-
tance, is clearly to be seen by comparing Figs 7.6 and 7.4.

7.3 A.C. operation


7.3.1 Resistor
With a sufficiently high mains frequency, and, in consequence, a sinusoidal
lamp voltage, there is little difference between d. c. and a. c. operation; the
same formulae apply (7.3), (7.4), (7.5), the current and voltage values then
being effective values.
At lower frequencies, such as 50 Hz, however, there is an important dif-
ference. As will be seen in Figs 7.7 and 7.8 the voltage over the resistor, and
thus also the current, is strongly distorted. The current is, indeed, zero over
a fairly large period f3 during each cycle. During this part of the cycle, very
little light is radiated and consequently the lamp flickers.
t This effect is particularly strong at high values of V 1 , since:
Vo
VI
f3 = 2 sin- 1 ---.
V 0 ]/2

Fig 7.7. Form of lamp voltage


and current when a fluorescent

L\ (\
lamp is operated on a.c. with
resistance in series. The hatched
areas are those in which the
supply voltage is higher than the
lamp voltage, i.e. current is
flowing through the lamp. Du-
ring the time {1 the current is
zero.
STABILISATION OF THE DISCHARGE ll3

a b
Fig. 7.8. Oscillograms of supply voltage and lamp voltage (a) and of lamp voltage and
current (b) when a fluorescent lamp is operated on a.c. with resistance in series.

The current distortion can be calculated with the help of equation 7.1. The
maximum value of the fundamental of the current is:
4
V0 V2-- V 1
n
Il = - - - - - -
R
whilst the nth harmonic has, as maximum value:
4 v,
In=-- (n = 3; 5; etc.),
nnR

so that with the ratio ~: = 05 usual in practice, the nth harmonic ex-
pressed as a percentage of the fundamental is:
In 82
- = -% (n = 3, 5, etc.). (7.6)t
It n

a b

Fig. 7.9. Oscillograms of supply voltage and lamp voltage (a) and of lamp voltage and
current (b) when a fluorescent lamp is operated on a.c. with a self-inductance in series.
The phase-displacement between lamp voltage and supply voltage is clearly to be seen.
114 FLUORESCENT LAMPS

7.3.2 Self-inductance (choke)


In the same way as a series resistor, the self-inductance must take up the
(in this case vectorial) difference between the no-load voltage and the lamp
voltage.
The product of this voltage difference and the lamp current is the apparent
power for which the choke must be designed. Since this power is purely
reactive for an ideal self-inductance, it does not lead to energy losses, so that
the efficiency of the circuit is much higher than in the case of a series-resistor.
The power factor, on the other hand, is lower. In practice, of course, certain
losses do appear, in the turns and in the iron core of the choke, but these
are small in comparison with the reactive power.
At sufficiently high frequencies, if the lamp voltage is sinusoidal, the lamp
current with a self-inductance as ballast is also sinusoidal. At lower frequen-
cies, such as 50 Hz, the lamp current remains, however, somewhat distorted.
Owing to the phase displacement which appears between mains voltage and
lamp current it is easily possible, nevertheless, to arrange that current flows
through the lamp during the whole of each half-cycle, so that the lamp flickers
either only very slightly or not at all. This thus constitutes a further important
advantage over the use of a series-resistor (Fig. 7.9).

t The calculation of the value of self-inductance required and the power


factor attainable is fairly complicated at low frequencies. We will give here
only the results calculated for a 'square-wave' lamp voltage and a loss free
linear self-inductance*.
The following voltage equation is obtained:
V0 2 = (11 xwL) 2 + (109 V1) 2 , (7.7)
whilst for the phase displacement between the zero values of mains voltage
and lamp current:
111 v,
COS((!=---,
Vo
and the power factor rx is no longer equal to cos q;, but:
v,
rx = 09- = 081 cos q;. (7.8)
Vo
This is a consequence of the current distortion which can be calculated
approximately, making use again of equation 7.1.
The maximum value of the fundamental is:

* W. ELENBAAS, et al., 'High pressure mercury vapour lamps and their applications',
Chapter 3, Philips Technical Library 1965.
STABILISATION OF THE DISCHARGE 115
and the maximum value of the nth harmonic:
4 v,
In= (n = 3; 5; etc.),
:n:n 2 wL

For V1 = 05 V0 the nth harmonic, expressed as a percentage of the fun-


damental is:
In 51
- = -% (n = 3; 5; etc.). (7.9)
Il nz
For the higher harmonics the amplitude decreases rapidly.

7.3.3 Capacitor
The use of a capacitor as ballast is, at first sight, ideal; its own electrical
losses can be kept very low, so that an efficiency can be obtained of almost
100%. The capacitor hardly gets warm, so that cooling presents no problems.
Finally, the chance of production of a disturbing hum is very small.
At sufficiently high frequencies these advantages can, indeed, be exploited.
At low frequencies, however, where the lamp voltage still has a square wave-
form, a capacitor is unusable owing to the distortion of the lamp current,
since the maximum value of the nth harmonic is:

and is, thus, independent of n. All uneven harmonics have the same ampli-
tude!
The unsuitability of a capacitor at low frequencies is also to be seen from
other considerations. The lamp voltage changes discontinuously each half-
cycle from + V1 into- V1 It can be assumed that, during the short time
occupied in this changeover, the mains voltage remains constan, so that the
whole voltage change must be taken up by the capacitor. According to the
equation
V=_gc
c

in which Vc is the voltage across the capacitor, Q the charge and C the capa-
city, Q must change at the same rate as Vc. A very large current surge is
needed to accomplish this rapid change in the charge. This surge passes
through the lamp each half-cycle, so that the light is radiated in flashes and
the life of the lamp seriously impaired (Fig. 7.10).
Summing up, the most attractive possibilities are:
d.c. - temperature dependent resistor.
a.c. - at low frequencies a choke, and at sufficiently high frequencies a
capacitor.
116 FLUORESCENT LAMPS

Fig. 7.I 0. Oscillogram of the lamp cur-


rent when a fluorescent lamp is operated
on a.c. with a capacitor in series. The
current consists mainly of high peaks of
short duration.

In many cases a different solution can offer special advantages. More com-
plicated circuits are then involved, as discussed in Chapter 8.

7.4 Consequences of current distortion


In the foregoing section special attention was given to current distortion
because of the undesirable consequences this can have, such as shortened
lamp life, fluctuations in light output, and overloading of cables and trans-
formers in the supply system.

7.4.1 Lamp life


The lamp wattage is determined, for a given lamp voltage, by the average
value of the lamp current. Normally, however, by lamp current is implied
the effective value of lamp current and this differs from the average value
to an extent determined by the current distortion. If the angular displace-
ment existing between the maximum values of the harmonics and that of
the fundamental is unfavourable (and this is usually the case), the desired
value of average current will be accompanied by a higher effective value and
a still higher maximum value of the lamp current. The lamp electrodes are
not designed for these high values, so that the lamp life is shorter in con-
sequence***.
Fig. 7.1 I shows how, for a specially made series of ballasts with different
distortion percentages, the effective value of the current depends on the ratio
of peak value to effective value (peak factor). The average value of the cur-
rent was kept approximately constant to give a constant value for the lamp
wattage. The resulting lamp currents for 5 of these ballasts are given in
Fig. 7.12.
The results of a series of life tests with 'TL' 40 W lamps on these ballasts
is shown in Fig. 7.13, in which the deleterious effect of the high currents is
clearly to be seen.

* R. N. THAYER and A. BROWNELL, 'Performance of Fluorescent Lamps on Leading


Power-factor Circuits', l/1. Eng., 44, 567- 570 (1949).
** E. F. LowRY, 'Thermionic Cathodes for Fluorescent Lamps and their Behaviour',
Ill. Eng., 46, 288-294 (1951).
STABILISATION OF THE DISCHARGE 117

y
Iz(%)

160 f-t-
I
I l,{

r"-
I

v
f--

140
V"
yL
I
f- f-

120 r-- r- i -
/
- t7 /
Fig. 7.11. Effective lamp current as
100 / a function of the distortion (expressed
1.5 2 2.5 3 in peak-factor) for a series of special-
Peak factor ly made ballasts.

a b

c d

Fig. 7.12. Oscillogram of lamp current for


5 of the ballasts of Fig. 7.1 I. Peak factor
values (from a to d) of 15, 165, 20 and
255 respectively.
118 FLUORESCENT LAMPS

L("'o)

-
700

,
50 ~

~ ....

20

10 -::-'---'--
' 'r- Fig. 7.13. Relative life of 'TL' 40 W
7.5 2 2.5 3 lamps, operated on the ballasts of Figs
Peak factor 7.ll and 7.12.

Agreement has been reached internationally that the ratio between the
maximum value of the lamp current and the effective value - the peak
factor- must not exceed 1.7, in order to avoid undue shortening of the
lamp life*.
It would have been more correct to have fixed the ratio between maximum
and average values, but this ratio is more difficult to determine.

7.4.2 Lamp flicker


When a high percentage of the uneven harmonics results in a lamp cur-
rent with high peak factor, the relative depth of the light ripple is larger and
stroboscopic effects become much more noticeable. At a mains frequency of
50 Hz the frequency of the ripple is 100Hz, which is so high for the human
eye that it is only directly visible in really bad cases.
The even harmonics in the lamp current are still more objectionable, the
ripple then having a 50 Hz component which is directly disturbing to the eye.
These harmonics can be limited by making the lamp perfectly symmetrical
and taking the necessary precautions in the design of the ballast. A useful
control measurement for this purpose is the difference in amplitude of two
successive half cycles of the lamp current.

7.4.3 Overloading of the supply system


If a three phase system is equally loaded with a series of identical electrical
appliances which only take the fundamental frequency, the neutral line carries
no current. If, on the other hand, they draw a distorted current from the
mains, the neutral does carry a current. The third harmonic is particularly
troublesome in this connection, since the components in the three phases

Ballasts for fluorescent lamps, Publications of the I.E.C., p. 23 (1962).


STABILISATION OF THE DISCHARGE 119
become additive in the neutral. If, thus, the appliances draw a current con-
taining 331-% of this harmonic, the neutral then carries the same current
(of three times the frequency) as the phases.
Since, in some mains, the neutral has a smaller cross-section than the phase
cables. this can become overloaded *.
In practice, the situation is usually much more favourable since a load com-
posed of identical appliances never occurs. There is usually a mixed load,
so that the third harmonics do not all have the same phase displacement
and a certain compensation is obtained. It is nevertheless desirable to keep
the current distortion, even per appliance, within bounds, for which purpose
the following formula has been agreed internationally: **
In IX
-~-----x (n = 3; 5; 7; etc.).
/1 10 (n- 267) 09

7.5 Measurement of ballasts


In order to see if a ballast supplies the correct power to the lamp and the
lamp current satisfies the various requirements concerning distortion, it is
necessary to load the ballast with a lamp of average characteristics. Other
alternative forms of load such as resistors or impedance networks usually
give misleading values of lamp power and do not allow an assessment of
the current distortion.
Now in order to select a lamp of average characteristics it is necessary to
have an average ballast, so that one is faced with a vicious circle. The problem
can best be solved by standardising a reference ballast with accurately defined
characteristics. For this purpose it is only necessary to make one series of
measurements with a large number oflamps in order to determine the correct
setting of the reference ballast. The reason why a ballast is to be preferred
to a lamp as primary standard lies in the fact that the properties of the ballast
are much easier to define than the properties of a lamp and, moreover, the
latter ones do not remain constant during the life of the lamp. International
agreement has been reached on the characteristics of reference inductive
ballasts for the various sizes of fluorescent lamps, the most important cha-
racteristics being a closely defined volt-ampere relationship, as linear as
possible, and accurately determined losses***. With the help of such a ballast
it is a simple task to select an approximately average lamp and, after calibra-
tion on the reference ballast, to use it to check a ballast under investigation.
The results of the two sets of measurements are compared and by use of
this relative method the lamp used need not have characteristics lying within
all too narrow tolerances.
Since the lamp voltage is strongly dependent on temperature, it is neces-
sary to keep the ambient temperature constant during measurement. A tem-
perature of 25 oc has been agreed for this purpose.
* E. STOLTE, 'Begrenzung des Stromoberwellengehaltes bei Entladungslampen', ETZ-B,
8, 22-25 (1956).
** 'Ballasts for fluorescent lamps', Publication 82 of the I.E. C., p. 23 (1962).
*** ibid, Annex 2.
Chapter 8

Lamp types and circuits


j. Funke and J. C. Moerkens

8.1 Introduction
The evolution of the fluorescent lamp and its applications has led to the
introduction of a large variety of lamp types and circuits. Since the charac-
teristics of the lamp are largely determined by those of the ballast, and vice
versa, it is desirable to deal with these two topics together.
The most important function of the ballast is to stabilise the lamp current,
since the arc voltage of most gas discharges decreases with increasing dis-
charge current (see Section 1.7). Further, the dimensions and dissipation of
the ballast are mainly determined by the desired power and starting proper-
ties of the lamp. The ignition voltage is of importance in this connection,
as the ignition voltage of the un-ionised gas mixture of a low pressure lamp
is appreciably higher than the normal arc voltage. The ignition voltage can
be quite considerably reduced by preheating the electrodes to emission tem-
perature (see Fig. 1.12). Most ignition systems therefore provide adequate
cathode heating as well as the required voltage across the lamp.
In the basic fluorescent lamp circuit, this pre-heating is realised with the
aid of a starter switch, which is in the closed position for this purpose.
Breaking of the contact leads to a voltage peak, which can be used to ignite
the lamp. Since this 'mechanical' starter contains a moving element, which
LAMP TYPES AND CIRCUITS 121
is subject to wear, many attempts have been made to avoid the use of this
starter, and to produce the necessary voltage and heating current by other
means. This has led to the development of a large number of starterless
circuits which, however, require special lamps. Various circuits of this type
will be discussed below.
Apart from its function in stabilising and starting the discharge, the ballast
should fulfil a number of other requirements, of which we will discuss the
following below:
(1) a high power factor, which ensures economic use of the supply system;
(2) a low percentage of harmonics in the current drawn from the mains (see
also Section 12.3);
(3) a high impedance for audio-frequencies, which are sometimes used for
signalling purposes over the supply system, is often required. Since prac-
tically always self-inductances are included in series with the lamp for
stabilisation purposes, most circuits have this high impedance automatic-
ally. However, special care must be taken in this connection when capa-
citances are shunted across the mains terminals; this is discussed in
Section 12.5;
(4) adequate suppression of radio interference caused by the lamp (see
Section 12.2);
(5) limitation of the luminous ripple in the light of the lamps, in order to
prevent stroboscopic effects.
The fluorescent lamp is essentially intended for a.c. operation, but d.c.
operation is possible with certain precautions and restrictions. This topic
was dealt with in Section 2.5.6.
Although fluorescent lamps are chiefly used for lighting purposes, it has
been found that the use of special fluorescent powders makes them also
suitable for other purposes, such as photocopying. These special purpose
lamps are discussed in Section 8.10.

8.2 Starter switches


The fundamental circuit for fluorescent lamp operation as shown in Fig.
8.1 contains a starter switch S. The purpose of this switch is to close the
circuit of the ballast and lamp electrodes, and to open it again after the elec-
trodes have been heated up sufficiently. The interruption of the current in

Fig. 8.1. Conventional method of start-


ing a fluorescent lamp. When the switch
Sis closed the short circuit current from
the choke flows through the electrodes.
The electrodes are thus heated and,
when the switch S is opened, the lamp
ignites by reason of the voltage surge
caused by sudden interruption in the
flow of current through the choke. 220Vrv
122 FLUORESCENT LAMPS

the choke provides a voltage surge across the lamp which makes it start.
The effectiveness of starting is to a large extent governed by the properties
of the starter switch.
According to their method of operation we may distinguish between the
following types of starter switches: manual switches, magnetic switches,
thermal switches, and glow switches. Only the latter two types have found
a wide-spread application, particularly the glow switch. We shall deal with
their characteristics in detail.
Manual switches are chiefly employed in desk lamps, magnetic switches
have never become popular, although some designs with interesting features
have been developed.
Right from the beginning it was found necessary to make the operation
of the starter switch automatic, i.e. that in case the first attempt to strike
the lamp is not successful, further attempts follow until the lamp has struck.
The present form of starter switches has been largely standardised. They
are housed in a small canister with 2 contacts for glow switches or with 4
contacts for thermal switches.
As a rule, a capacitor of 0006-002 [J.F is built into the canister and con-
nected in parallel across the lamp and starter. This capacitor suppresses to a
large extent the radio interference which might be caused by the fluorescent
lamp.
The following requirements must be met by a good starter:
(1) The contact should remain closed long enough to permit a sufficient
heating of the cathodes;
(2) The surge should be high enough to ensure starting under adverse con-
ditions (under-voltage, low temperature);
(3) It should be capable of performing many thousands of operations.
(4) It should not close at the lamp voltage, even under adverse conditions.

8.2.1 Glow switches (see Fig. 8.2)


The glow switch consists of a glass envelope filled with rare gas and con-
tains two electrodes, of which at least one consists of a bi-metallic strip.
This strip, composed of two thin layers of metal, having different coefficients
of expansion, has the property of bending when heated. When a sufficiently
high voltage, here the mains voltage, is applied to the switch, a glow dis-
charge takes place in the rare gas and the heat of the cathode fall causes the
bimetallic strip to deflect, thereby closing the contacts, so that the lamp
electrodes are heated. The glow discharge then ceases since the starter is
short circuited, the bimetallic strip cools off and opens the contacts. This
action is repeated a few times, so that the cathode can reach its required
emission temperature. (See Chapter 5, Fig. 5.1.) The voltage surge, resulting
from interruption of the current by the starter, then ignites the lamp. This
surge should not be too high, since otherwise the lamp would strike before
the cathodes are properly heated, with corresponding damage.
A very important characteristic of the glow switch is its closing voltage.
This voltage must lie between two limits; the upper one is given by the
LAMP TYPES AND CIRCUITS 123

PHil:
'/C

Fig. 8.2. Glow switch starter.

lowest voltage where the switch is expected to operate, the lower one is
governed by the highest value which the arc voltage of the lamp can reach.
Should the starter switch close again under the influence of the lamp voltage
then the lamp would be short circuited and start blinking off and on. Since
the open circuit voltage of the ballast is in most cases roughly twice the arc
voltage of the lamp, it is evident that for lamp types with widely different
arc voltages different starter switches have to be used.
Apart from this requirement there are other factors governing the design
of the starter switches. We will mention the requirement of a relatively small
closing time, depending amongst other things on the properties of the bi-
metal, on the gap width and on the speed of heating up of the bimetal, which
in its turn also depends on the glow current. The latter is to a certain extent
determined by the impedance of the ballast. The total time that the switch
is closed, the preheat current and the dimensions of the electrodes determine
the temperature of the electrodes at the moment of striking. The preheat
current provided by the ballast is also of importance with regard to the
number of operations the starter switch can withstand. These aspects have
led to the development of a series of starters, each destined for one or more
particular lamp types.
A special execution of the glow switches is represented by the safety
starters ('watch dog type'), Since a glow switch will continue its efforts to
strike a lamp, which is defective but has its electrodes still intact, these
efforts may cause annoying light flashes through the lamp. Apart from being
troublesome, this action will reduce the life of the starter and may lead to
overheating of ballasts of poor design. It can be avoided by incorporating
a cut-out in the starter.
With the present long life of fluorescent lamps the occurrence of starters
trying to start lamps with deactivated electrodes has diminished consider-
ably. This has made the 'watch dog' starter more or less obsolete.
124 FLUORESCENT LAMPS

8.2.2 Thermal starter switches


In contrast with glow switches, of which the contacts are normally open,
the thermal switch is closed at the beginning of the starting operation (see
Fig. 8.3).

Fig. 8.3. Circuit diagram for use with


thermal starter switch.

Here again, contact is made and broken by a bimetallic strip. When the
current is switched on, it passes through the cathodes of the lamp, as well
as through the heater H in the switch, thus providing a very good pre-heating
of the cathodes; the heater H heats the bimetallic strip which deflects and
breaks the contact, and, when the lamp has ignited, current continues to
flow through H; the heat thus produced is sufficient to keep the contacts
open. A disadvantage of this type of switch is that, when the lamp is ex-
tinguished, the switch is not immediately ready to operate again, because the
thermal inertia involves a certain amount of time for closing the contacts.
The consumption of power by the heater H whilst the lamp is burning may
be regarded as a minor drawback.
This type of starter switch is excellent from the point of view of ignition
but it has never become popular, owing to the fact that it necessitates four
connections and thus further complicates the wiring of the lamp. Its excellent
qualities arise from the good pre-heating of the cathodes and from the fact
that the contact opens in vacuum or in air (depending on the design) which
results in a sudden interruption of the current, and thus in a higher voltage
surge. In the glow switch the low pressure gas filling invariably involves a
certain loss in power, at the moment of opening, thus rendering the voltage
surge less effective.
The high voltage surge of the thermal switch does not constitute a danger
to the cathodes, since in this case they are properly pre-heated.
Apart from the simple type of thermal switch described above, some manu-
facturers have designed more complicated devices. These contain as a rule
more than one bimetal strip and a few auxiliary contacts. The result is a
starter which gives an appropriate preheat of the cathodes as well as a high
ignition kick. This results in very good starting properties at low tempera-
tures as well as in long lamp lives.
LAMP TYPES AND CIRCUITS 125
8.3 Lamp types for switch start operation
Ever since fluorescent lamps were first introduced there has been a steady
growth in the number of types. The reasons for this are not far to seek, since
once the first models of relatively low power had become established, interest
was aroused in larger units, and this was in due course followed by a demand
for lamps for special purposes.
In view of the fact that the negative characteristic of gas-discharge lamps
renders the use of a ballast unavoidable, it is not possible to increase the
arc voltage beyond about 50% of the applied voltage if a stable discharge
is to be maintained (see Section 8.4). In the United States the standard mains
voltage is 120 V; therefore the original models of fluorescent lamps were
designed for an arc voltage of some 60 V. These were lamps of 15 and 20 W,
in lengths of 18 and 24 in respectively. For use on 220-240 volts supplies
the lengths of the tubes were doubled to 36 and 48 in, and the wattage to
30 and 40 W. These tubes were manufactured in diameter of 1 and 1t in
respectively.
In addition to these American types, Philips introduced a lamp of 100 em
length for 25 W and a lamp of 60 in for 65 W. All these lamps are equipped
with standardised bi-pin caps.
Development in England followed entirely different lines; by 1939 an
80 W lamp had been produced in a length of 5 ft and a diameter of 1t in.
Owing to the contingencies of the war, this was the only fluorescent lamp
on which production was concentrated. For the sake of rapid introduction
it was provided with two standard bayonet lamp caps, and to this day this
lamp is still being used in England, although after the war many American
types were adopted. Today 80 W lamps are made chiefly with hi-pin caps.
In recent years 65 W lamps are getting a fairly large share of the market,
moreover a 65/80 W lamp has been developed. This lamp can be used on
either 65 or 80 W ballasts.

Table 8.1 Data of switch start lamps


approx.
nominal nominal diameter arc nominal luminous flux
watts length voltage current (Cool White)

I5 I8 in I in 54 033 A 900 lm
20 24 in It in 57 037 A II50 lm
25 IOO em It in 94 029 A I 800 Im
30 30 in I in 96 0365 A 2 200 lm
30 30 in 11-in I 8I 0405 A 2 050 lm
40 IOO em H in 80 056 A 2 850 lm
40 48 in It in I03 043 A 3 200 lm
65 60 in It in 110 067 A 4 900 lm
80 60 in lj-in 99 087 A 5 700 lm
I25 96 in It in I49 094 A 9 500 lm

Besides the lamp types mentioned so far, all of which are made in a dia-
meter of 1 in or 1-!- in, a 90 W lamp, 5 ft long and 2t in diameter is used in
the United States. Owing to the low arc voltage (about 60 V), this lamp is
126 FLUORESCENT LAMPS

not an economic proposition on 220 V supplies, since the ratio of arc voltage
to supply voltage is unfavourable. It is employed mainly for illuminating
large factories. The advent of the new highly loaded lamps (see Section 8.5)
has rendered the 90 W lamp more or less obsolete. Table 8.1 gives a survey
of the more important data of the principal lamp types.
The luminous flux is given for the cool white colour. For other colours
there is a fairly constant relationship between the luminous flux for those
colours and standard cool white. Assuming the latter to be I 00 %, the per-
centage for other Philips colours is approximately as shown in Table 8.2.

Table 8.2
colour luminous flux
colour designation
number (% of standard cool white)

29 standard warm white 100


54 standard daylight 85
32 de luxe warm white 65
34 de luxe white 70
55 colour matching 65
27 softone 55
37 trucolour 55
57 daylight de luxe 60
(artificial daylight)

Apart from the above mentioned types a number of lamps in i in dia-


meter have been developed. Their principal data are given in Table 8.3.

Table 8.3 Data of 5/8 in lamps


approx.
lamp type nominal diameter arc nominal luminous flux
length voltage current (Cool White)

4W 6 in tin 30 015 A 150 lm


6W 9 in tin 45 0155 A 300 lm
8W 12 in tin 58 0165 A 470 lm
13W 21 in fin 98 017 A 940 lm

These lamps are particularly suited for use in those cases where only
limited space is available.
Another solution to the problem of the sometimes impractical dimensions
of normal fluorescent lamps is the circular lamp (Philips 'TL'E type). Three
types have been put on the market. Their principal data are listed in Table
8.4.
LAMP TYPES AND CIRCUITS 127
Table 8.4 Data of circline lamps

la mp type nominal approx. luminous


outer diameter arc voltage
current flux (Cool White)

22 w 8t in 62 v 039 A 1100 lm
32 w 12 in 84 v 043 A 2 000 lm
40W 16 in l!OV 042 A 2 850 lm

Apart from the circline lamps a few types of U-shaped lamps have been
developed. Some data of the most important types are given below in Table
8.5.
Table 8.5
approx.
lamp type tube distance arc voltage lamp luminous flux
diameter between legs current (Cool White)

16W
20W
26
26
mm
mm
56 mm
92mm
94
63 I 020
035
A
A
850
1100
lm
lm

I
40W 36 mm 92 mm 114 041 A 2 900 lm
65 w 38 mm 92 mm 125 062 A 4 800 lm

In Fig. 8.4 examples of circline and U-shaped lamps are shown.

Fig. 8.4. Circline and U-shaped lamps.


128 FLUORESCENT LAMPS

8.4 Starter circuits


Starter circuits can realise the necessary stabilisation of the gas discharge
in two ways;
(1) inductively (by means of a choke coil or an auto leakage transformer);
(2) capacitively (with a capacitor in series with a choke, or a capacitively
loaded leakage transformer, or on the sequence start principle).
Now when a reactance is placed in series with the lamp, the power factor
will be low. It can be raised in a number of ways:
(1) by means of a dual lamp circuit, which combines inductive and capacitive
stabilisation;
(2) by shunting a capacitor across the mains terminals in an inductive cir-
cuit;
(3) by increasing the (inductive) magnetising current of a transformer, when
capacitive stabilisation is used; this gives e.g. a sequence-start circuit.

8.4.1 Inductive stabilisation


The fact that the gas discharge in a fluorescent lamp has to be stabilised
(see Chapter 7) means that these lamps are more suitable for a.c. than for
d.c. operation, since a choke coil provides a means of bridging the gap be-
tween the supply voltage and the lamp voltage with the minimum loss of
energy.
If an inductive ballast is used, it would seem to be an obvious idea to
design the circuit so that the lamp gets the maximum share of the supply
voltage, and the choke the minimum; with a given lamp power, this then
gives the smallest possible ballast. However, there are a number of consi-
derations which modify this conclusion in practice:
(a) If we regard the lamp as an ohmic resistance, we find that the division
of the supply voltage between the lamp and the ballast is as given in Fig.
8.5. The lamp current I 1 is in phase with the lamp voltage V 1 ; the magnitude
of this current is given by
Vm-""Vl _ VL
wL--wL
where VL is the voltage across the choke, and Lis its self-inductance. Now
if the vector Vm, representing the supply voltage, falls e.g. to Vm' as a result
of mains fluctuations, then VL will also fall to VL', since the magnitude of
the lamp voltage remains constant to a first approximation.

Fig. 8.5. Vector diagram of an ohmic resistance


and a self-inductance connected across the mains.
LAMP TYPES AND CIRCUITS 129
It will be clear that the relative decrease in VL under these conditions will
be greater than that in Vm, and since the lamp current / 1 is directly propor-
tional to VL, the relative decrease in the lamp current (and hence also in
the lamp power) will also be greater than the decrease in vm causing it.
This effect will be greater when the nominal value of V1 is closer to Vm.
In circuits where Vm V" we may take VL R:; Vm to a first approximation;
a change in Vm will then be accompanied by an equal change in VL, and
hence in the lamp current. The optimum situation, where the light output
of the lamp is proportional to the mains voltage, can thus only be realised
when the lamp voltage is small compared with the mains voltage, in other
words when the ballast is relatively large. It has been found in practice that
Vm/V1 R:; 2 represents a good compromise for fluorescent lamps; a change
of 10% in the mains voltage then gives a change of about 20% in the luminous
flux. This ensures that the lamp will burn fairly steadily under normal con-
ditions; see also Fig. 8.6.

% 100

t
--
90
Luminous
flux('{)) 80
I
I t ~ ~
e--
- f--- ~~
current(I1) I~It
70 I /)

j/

-vf
I
60

50

~0

30
~ ~ u ~ ~ ~ n ll ~ u
---I~~Vml
i'Vf
Fig. 8.6. Lamp current and luminous flux at a mains voltage 10% below the nominal value,
as a function of the ratio Vm/V1 The luminous flux and lamp current at the nominal voltage
are taken as 100 %.

(b) If we plot the quantities of Fig. 8.5 as functions of time, we get Fig. 8.7.
At time t when the lamp current goes through 0, the lamp goes out, and
0 ,

cannot be re-ignited in the opposite direction until the available voltage equals
or exceeds the arc voltage of the discharge; in other words, the instantaneous
value of the mains voltage at time t0 must be greater than V 1 However, as
vm falls the phase angle q; also decreases, so that the proportion of the mains
voltage available for re-ignition falls off extra quickly.
The situation is illustrated for various mains voltages in Fig. 8.8. In Fig.
8.8c, the mains voltage available at tv is just sufficient. Taking both the lamp
130 FLUORESCENT LAMPS

Fig. 8.7. The quantities of Fig. 8.4 plotted as functions of time. The instantaneous value
of the mains voltage at time t 0 must be at least equal to the lamp voltage if the lamp is
to be re-ignited.

voltage and the supply voltage fluctuations into consideration, we may


conclude that Vm/ V1 should be at least about 19 if we are to avoid sudden
extinguishing of the lamp.
(c) As explained in Section 8.2, the starter switch must not close while the
lamp is burning; but when the lamp is not ignited, the starter switch must
close reliably and rapidly.
Taking the above three points into consideration, we may conclude that
with inductively stabilised lamps the ratio VmlV 1 should be at least 20.

a b

Fig. 8.8. As the ratio Vm / V1 falls, the


voltage available for re-ignition at the
end of each half-period also falls.
LAMP TYPES AND CIRCUITS 131
Now the mains voltages most commonly found in practice lie either be-
tween 200 and 250 V or between 110 and 127 V. Inspection of the list of
arc voltages of the most popular lamps (see Section 8.3) then shows that
the following combinations of lamps can be used in series with a single self-
inductance:
With Vm in the range 200-250 V:
choke + 1 lamp up to 80 W (Fig. 8.9a)
or choke + 2 lamps in series, up to 20 W each (Fig. 8.9b).
With Vm in the range 110-127 V:
choke + 1 lamp up to 20 W (Fig. 8.9a).

a b
Fig. 8.9. Lamp(s) stabilised by a single self-inductance; such a circuit must satisfy the con-
dition Vm/V1 ;;;:. 20.

By way of example, let us consider what happens when one 40 W fluores-


cent lamp is used at Vm = 220 V and 50 Hz.
Operating data:
Wballast = 10 W VAchoke = 74 VA
Dimensions of ballast: 38 x47 x 108 mm
Overall cos q; = 051
V1 = 103 v, 11 = 440 rnA, VL = 168 v.
During ignition, the choke will receive the full 220 V, because of the low
resistance of the cathodes. The current under these conditions will be higher
than that during normal operation, being of the order of 650 rnA, so that
the cathodes will warm up quickly.
However, this simple circuit with one choke cannot be used when the arc
voltage is chosen nearer the mains voltage, i.e. when the ratio Vm/V1 is
appreciably smaller than the optimum value of 2 mentioned above. In this
case, one must use some kind of step-up transformer; this transformer does
not generally need to have separate primary and secondary windings, i.e.
it can be made as an autotransformer (Fig. 8.10). The corresponding vector
diagram is shown in Fig. 8.11. The line 1-2 represents the mains voltage,
and 1-3 the secondary voltage, which is equal to the sum of 1-4 (the lamp

Fig. 8.10. If Vm 2 v, a step-up transformer should be


used.
Fig. 8.11. (a) Vector diagram for the circuit of Fig. 8.9. The windings 1-2 and 2-3 are
tightly magnetically coupled.
(b) If the windings 1-2 and 2-3 are not so tightly coupled, we get a spread impedance
which gives rise to the voltage vector 3-3'.
(c) Vector diagram of an auto leakage transformer as shown in Fig. 8.12. The separate
self-inductance of Fig. 8.10 is now no longer necessary.

voltage) and 4-3 (the choke voltage). In most cases where a transformer is
used, one tries to make the magnetic coupling between the two windings as
high as possible, so that the secondary voltage 1-3 of the unloaded circuit
is the same as that under full load. In the ballasts for fluorescent lamps,
which are usually long and thin in shape, this is achieved by having one
winding over the other. If however the two windings are placed side by side
(Fig. 8.12), we get a certain magnetic spread, i.e. not all the lines of force
of the secondary windings pass through the first winding, and vice versa.
The 'leakage lines' play no role in the transformer effect (magnetic coupling),
but form separate self-inductances which can be regarded as being in series
with the secondary winding. The voltage across this 'spread self-inductance'
is represented by V3 _ 3 ' in Fig. 8.11b, a vector which is at right angles to the
current vector. As a result of this effect, the voltage vector 2-3 (unloaded
state) rotates to 2-3', which is the resultant of 2-3 (due to the transformer
effect) and 3-3' (due to the spread self-inductance). The separate choke 3--4
can then be made smaller (3' --4). This principle can be carried further by the
use of magnetic shunts (Fig. 8.12), which allow the spread to be made so
great that the resultant self-inductance can completely replace the choke.

u------
'
--- ~ or --------~1
l ' .......
: pnm sec

Il ____ ------~ o l __________ I Fig. 8.12. Structure of a matching leak trans-


former. The magnetic spread between the two
windings can be adjusted by means of the two
s iron 'shunts' S.
LAMP TYPES AND CIRCUITS 133
This gives the matching auto leakage transformer, widely used in this form,
whose vector diagram is shown in Fig. 8.11c; the secondary voltage vector
here rotates from 2-3 in the unloaded state to 2-4 at full load.
In many cases, the vector 1-2 ~ 2-3 (e.g. when we have step-up from
110 V to 220 V). The length of l-3 is then representative of the total amount
of VA in the transformer. Since 3-4 is the corresponding measure of the
VA which would be consumed in the self-inductance if the mains voltage
was l-3 (cf. the vector VL in Fig. 8.5), the ratio l-3/3-4 gives an impres-
sion of how expensive an auto leakage transformer is going to be as a ballast.
The use of such a transformer as ballast allows:
8 ft lamps to be used at mains voltages of 200-250 V and 4-5 ft (25-120
W) lamps to be used at mains voltages of 110-127 V.
Example: A 110 V and 50 Hz auto leakage transformer for one 40 W
fluorescent lamp has :
100 VA Wballast = 13 W COS cp ~ 055
External dimensions: 38 x47 x 153 mm
Open circuit voltage = 206 V.
We then have V1 = 103 V, and I 1 = 440 rnA.

8.4.2 Parallel compensation


The use of a self-inductance in series with the lamp to give the desired low
dissipation in the ballast entails the disadvantage that the overall cos cp
of the circuit will be low (see e.g. the phase shift cp in Fig. 8.5). The value
of cos cp can be increased by shunting a capacitor across the mains ter-
minals.
As will be seen from the vector diagram of Fig. 8.13, the phase difference
between the lamp current and the mains voltage is cp', while the current
through the power-factor capacitor, Ic, is 90 out of phase with the mains
voltage. The sum of the currents I, and Ic is equal to the mains current,
which shows only a small phase displacement cp with respect to the mains
voltage. The value of cp is governed by that of the capacitance C, as shown

Fig. 8.13. Vector diagram of current and voltage for a fluorescent lamp with choke ballast.
V1 = arc voltage, I 1 = lamp current; I c = current in power-factor capacitor; V L = voltage
across choke; Vm = mains voltage; Im = mains current.
134 FLUORESCENT LAMPS

cos~

/
v ~ _______.. 1. 0

0.9
/
/ .............
1,00 .......

350
~ v 0. 8

0.7

---
300 fl. 6

250
/ ...............
.......__ ........ -0. 5
_______..
2 3 5 6
Fig. 8.14. Mains current and power factor for a 40 W lamp with choke ballast on the
220 V mains, as functions of the capacitance in the circuit.

in Fig. 8.14 which gives the power-factor correction curve for a 40W lamp
with choke ballast, together with a curve representing the mains current as
a function of the capacitance. As is to be expected, the minimum mains
current occurs at practically the same value of C as the maximum power
factor which is usually between 095 and 098, not unity as might be expected.
This slight reduction of the maximum power factor is due to the fact that
the current flowing in the lamp is not exactly sinusoidal, but contains a
certain percentage of odd harmonics which cannot be compensated by the
capacitor, so that absolute compensation is not possible.
It is sometimes possible to correct the power factor of a group of n lamps
from a central point, using a capacitance equal to n times that needed for
a single lamp.
These simple methods of power-factor correction are not always feasible.
There has been a growing tendency of recent years to use mains operated
electrical signalling devices in which voltages of frequencies from 175 to
1 500 Hz are superimposed on the 50 Hz mains voltage to actuate relays,
which then execute a variety of switching operations ('centralised ripple
control'). Capacitors across the mains do not offer much impedance to such
audio frequencies, and therefore draw a considerable amount of power from
the signalling system. This point is discussed in further detail in Chapter 12.
Another point which must be taken into consideration when installing
circuits with capacitors shunted directly across the mains is the choice of
the switch used.
When the circuit is switched on, the instantaneous mains voltage is applied
directly to the uncharged capacitor. Depending on the moment of switching
on and the presence of other (parasitic) impedances in the circuit, this can
give rise to large current peaks. The situation is aggravated by the contact
bounce from which every switch suffers to a certain extent (i.e., the fact that
the current is cut off a number of times before the contact is permanently
LAMP TYPES AND CIRCUITS 135
made). As a result of this effect, the switch must be designed for at least
twice the current that would otherwise be necessary.
In view of the above points, it is understandable that many users prefer
to employ dual-lamp circuits, which combine inductive and capacitive
stabilisation.

8.4.3 Capacitil'e stabilisation

L-C series circuit


As was shown in Chapter 7, the lamp voltage contains a large proportion
of higher harmonics. Since the impedance of a capacitor decreases in inverse
proportion to the frequency (w occurring in the denominator), it will be
clear that if only a capacitance is placed in series with the lamp, the lamp
current will be completely distorted. One well-known method of suppressing
the higher harmonics in the lamp current is to include a self-inductance in
the circuit in series with the capacitance. The following points are of great
importance in this connection.
1. The fluctuations in lighting level caused by mains fluctuations with this
system are generally less than with purely inductive circuits. In general,
a I0% change in the mains voltage may be expected to give rise to a
change of about I 0% in the lighting level, which can be regarded as a
very good result*. This means that the ratio Vm/V1 can safely be reduced
to 13-14.
2. The introduction of a capacitor into the circuit makes the vector diagram
more complicated than in Fig. 8.5; see Fig. 8.15.
When the lamp current and the arc voltage fall to zero in this circuit
at the end of each half cycle (time t0 ), we may write Vm = VL + Vc.

-----
+:,,
--
v,~:
Lto :I1 '

~~~----~r---~~
to \! ~ It
,,,,
Fig. 8.15. Vector diagram of a self-inductance (L) V,~!
"
and a capacitance (C) in series with an ohmic mto ::,,
resistance. The voltage across the capacitance has
been chosen greater than that across the self-in-
ductance, so that the lamp current (/1) is leading
with respect to the mains voltage (Vm). "
* The interested reader is referred to Section 3.8 of the book on High Pressure Mercury
Vapour Lamps and their Applications by W. ELENBAAS published by Philips Technical
Library, for further details.
136 FLUORESCENT LAMPS

The instantaneous values of these quantities can be found by projecting


the corresponding vectors on to the time axis t 0 ; the projections of V 1
and / 1 under these circumstances will be zero, of course.
Since not enough voltage is now available to re-ignite the lamp, it will
be quenched for the moment, and the lamp current will remain zero.
However, this means that the choke voltage VL must fall off to zero, as
a result of which the voltage across the lamp will increase by the same
amount, thus allowing the lamp to re-ignite. Unlike the case with purely
inductively stabilisation, therefore, the reliability of re-ginition after each
half cycle does not depend directly on the mains voltage, but on the choke
voltage. This means that a minimum value of Vm/V1 is no longer required.
3. In connection with the closing of a glow switch starter, the voltage across
this switch before the lamp is ignited must be about twice the burning
voltage of the lamp, as is also the case with inductive circuits.
Capacitive glow switch starter circuits thus allow the use of the same lamp
combinations as purely inductive circuits in the different voltage ranges,
namely:
for 200-250 V: choke+ cap.+ 1lamp "'( 80 W (Fig. 8.16a)
choke+ cap.+ 2 lamps"'( 20 W (Fig. 8.16b)
for 110-127 V: choke+ cap.+ 1lamp "'( 20 W (Fig. 8.16a).

a b
Fig. 8.16. Stabilisation by means of a choke and a capacitor in series. The condition
Vm/V1 > 20 is now only necessary in connection with the glow-switch starter, not be-
cause of the stability of the discharge with a varying mains voltage.

The advantages mentioned under (1) and (2) above cannot thus be com-
bined with the advantage of cheaper ballasts.
It may be remarked that if thermal switches are used for starting the lamps,
the disadvantage mentioned under (3) is eliminated; it is then possible to
use 8 ft lamps with arc voltages of up to 180 V in the mains voltage range
from 220 to 250 V (see Fig. 8.16a). However, as the thermal switch is not
very popular, this combination is not much used.
If dual lamp circuits are used, we only need one capacitor per two lamps,
and the self-inductances present are sufficient to deal with the high fre-
quency signals applied to the mains.

Capacitively loaded auto leakage transformer


When the mains voltage is less than 200 V and it is still desired to use long
fluorescent lamps, the obvious solution is to use a step-up transformer (Fig.
8.17). Here again, the transformer and the self-inductance can be combined
in the form of an auto leakage transformer.
LAMP TYPES AND CIRCUITS 137

Fig. 8.17. If VmfV1 < 2, a step-up transformer can


be used.

However, a complication arises here. If we compare Fig. 8.11c with Fig.


8.18, which gives the vector diagram of a capacitively loaded auto leakage
transformer, we see that as a result of the capacitive stabilisation the lamp
current no longer has a phase lag of cp with respect to the mains voltage, but
a phase lead of cp. Consequently, the voltage vector of the self-inductance,
which is perpendicular to the lamp current vector, is also in the other direc-
tion.
If now we introduce shunts to replace the separate choke by a spread self-
inductance, we find that the voltage 2-3 in the unloaded state rises to 2-4
in operation, thus giving a sizable increase in inductance. In the case of
Fig. 8.llc, on the other hand, we saw that the vector sum of 2-3 and 3-3'
gave the vector 2-4, which is equal to 2-3 but in a different direction.
This increase in inductance gives rise to an impermissibly high distortion
of the lamp current. Various methods have been devised to get round this
difficulty, but the final price of the circuit- e.g. for a single lamp with an
arc voltage of 100-110 V and a mains voltage of the same order of magni-
tude - is so high that these systems have never become popular.

'5
\
I

'' '
I
I

,,/'/
Fig. 8.18. Vector diagram of a capacitively loaded auto leakage /
/

transformer. The operating voltage across the secondary winding


(2--4) is much greater than the unloaded voltage (2-3), which
gives rise to increased induction and hence to distortion of the
lamp current.

The sequence start circuit with starters


We have seen in our discussion of capacitive stabilisation that once the
lamp has been ignited the ratio Vm/V1 can safely be made a good deal lower
than 2, thus allowing a closer approximation to the ideal solution of the
smallest possible ballast for a given lamp power. It is only for the operation
of the starter that this ratio often has to be made about 2.
138 FLUORESCENT LAMPS

Fig. 8.19. Principle of a sequence start circuit in its


simplest form, with a separate transformer to
which the coil, capacitor and lamps are connected.

This difficulty can be got round by using two lamps in series per capacitive
ballast (see Fig. 8.19). The open circuit voltage of the transformer (if needed)
only exceeds the sum of the two arc voltages by enough to ensure steady
burning of the lamps even with maximum negative fluctuations in the mains
voltage. A value of V0 /V1 = 13-14 will then suffice. This open circuit
voltage is not enough to actuate both starters, but it will actuate one, in
series with the auxiliary capacitance shown in Fig. 8.19. Once the lamp in
question has been started, nearly the whole open circuit voltage will be
applied across the second starter, which will thus be actuated in its turn.
In view of its principle of operation, such a circuit is known as a sequence
start circuit.
Of course, the capacitive stabilisation means that the power factor has a
phase lead, but as a result of the more favourable V0 /V1 ratio this factor is
quite high (about 07) anyway. The inductive magnetising current which is
automatically present at zero load is increased by the provision of an air
gap in the iron circuit, thus making the overall cos ([! quite high (;;:, 090).
This principle can be used e.g. with a mains voltage of 200-250 V and two
lamps with arc voltages of about 110 V in series.
The sequence start circuit is in principle somewhat cheaper than the dual-
lamp circuit, and has somewhat lower losses. On the other hand, the two
lamps are not independent of one another, because of the series arrange-
ment. This circuit is not much used in this form, but it is quite widely used
with starterless lamps (see Section 8.6).
However, the sequence start principle is much more attractive than the
use of two separate circuits for two lamps where the mains voltage is ap-
preciably lower than the sum of the arc voltages. This is the case with the
combinations of two 25-80 W lamps at a mains voltage of 110-127 V, or
two 8 ft lamps at a mains voltage of 200-250 V.
It will be seen from inspection of Fig. 8.19 that this change from a high
to a low mains voltage means that the step-up transformer 1-2-3 has to be
made appreciably larger to give the same open-circuit voltage.
With the aid of this larger transformer, we can give the supply voltage
for the lamps such a distortion that it has the same content of higher har-
monics as that produced by the lamps themselves, but in anti-phase with
the latter. The method for suppressing the higher harmonics by including
a self-inductance in the circuit is thus modified by having an anti-phase
voltage produced by a transformer, so that the extra choke in Fig. 8.17 can
be left out in part or completely.
LAMP TYPES AND CIRCUITS 139

Fig. 8.20. Principle and construction of


a transformer for a sequence start cir-
cuit which does away with the need for
a separate self-inductance for suppress-
ing the higher harmonics.

A transformer of this type is shown in Fig. 8.20. It differs from a normal


auto leakage transformer in the presence of an air gap in the secondary part
of the core; however, this air gap does not extend right through the cross-
section of the iron. Furthermore, the natural magnetic spread between the
windings is generally so large that extra magnetic shunts are unnecessary.
If the above-mentioned air gap were not present, the various magnetic fields
and voltages in the system would vary with time as shown in Fig. 8.2la,
where <Ps 1 and <Ps 1 H add up to the total field </Jprim and Vs 1 and Vs 1 H add
up to the mains voltage vm. It goes without saying that there is a phase
shift of 90 between the voltages and the fields .
If now a partial air gap is provided in the secondary part of the iron cir-
cuit, then at low inductions (when <Ps 1 is small), where the parts marked K
in Fig. 8.20 are not saturated, the voltages and fields will remain as shown
in Fig. 8.2la. This will be the case between the times 11 and t 2 in Fig. 8.2lb;
the form of the curves here will thus be the same as for the same period in
Fig. 8.2la. However, if <Ps 1 increases so far that the regions K are saturated
(which occurs at time t 2 ), then <Ps 1 cannot increase any further. Since how-
ever </Jprim continues to vary sinusoidally, with Vm, <Ps 1 H will increase very
sharply after t 2 , as shown in Fig. 8.21 b.
Since V = - d</J fdt, the induced voltage Vs 1 will be zero when the field
<Ps 1 is constant, while Vs 1 H on the other hand will increase sharply under
these conditions. At any given moment, of course, the relation Vs 1 + Vs 1 H
= Vm remains satisfied.

pnm
prim.
vm
s,
vs,

....
~

------
t, 12
a b
Fig. 8.2 I. (a) The magnetic fields and voltages for the transformer of Fig. 8.20 in the ab-
sence of the partial air gap in the iron core.
(b) The provision of the air gap in Fig. 8.20 alters the magnetic fields and voltages.
140 FLUORESCENT LAMPS

~ec.tot.

Fig. 8.22. Secondary voltage of auto leakage trans-


former with partial air gap. This secondary voltage
is equal to the sum of the sinusoidal mains voltage and
the distorted voltage across the secondary winding,
which are also shown in the figure.

Since S 1 and S 2 are closely coupled, the same sharp peak will occur in
S 2 as was found in S 1 The total secondary voltage (Vsec tot), which is equal
to the sum Vm +
V82 , will thus be distorted as shown in Fig. 8.22. The degree
of distortion depends on the size of the regions K compared with the total
width of the core, and to a lesser extent on the primary induction. Fig. 8.23
gives the percentage distortion (the proportion of higher harmonics as a
percentage of the amplitude of the fundamental) as a function of the depth
to which the core is sawed out, for three values of the induction. Now while
we can control the percentage of harmonics quite well by varying the size
of the air gap, it is more difficult to obtain just the desired amplitude of
harmonics - and it is this amplitude which should be roughly equal to the
amplitude of the distortion produced in the lamp.
If the lamp to be used (and hence the magnitude of the square-wave voltage
%
Distortion

t 30

20

17000gauss)
14000 gauss :tdharmonic
10 -----.......,.,12o"="o'"'"og=-a-uss

sth harmonic
17000gaus~ 7th harmonic

0 10 20 30 40 50 60 70 80 90 100
-~ Bridging of air gap (x/a in%)
Fig. 8.23.
Percentage distortion of the fundamental, as a function of the depth to which the air gap is cut
choke data: transformer data:
Spr = Ssec =
1150 w- 035 0 a= 14.7 mm
gap width = 15 mm b = 6.5 mm
c = 12.3 mm
d = 2.5 mm
LAMP TYPES AND CIRCUITS 141
in it) is known, we can calculate the amplitude (in volts) of the 3rd, 5th etc.
harmonics produced by the lamp. The transformer must then produce at
least the same amount of harmonics. If the size of the air gap has been
correctly chosen, we know the percentage of higher harmonics with respect
to the fundamental, but the amplitude of the higher harmonics, in volts,
can still be too small. It can then be brought up to the desired value by
varying the fundamental (i.e. the 50 Hz component) too. This means that
the secondary winding should have a minimum (50 Hz) voltage. Now if for
example the difference between the mains voltage and the open circuit voltage
is small, so that a small secondary winding would be sufficient as regards
power production, such a small secondary winding would not produce
enough distortion. The secondary winding must thus be chosen bigger, so
that either the open circuit voltage is bigger too, or a lower tapping must be
taken from the primary coil.
Apart from the amplitude of the harmonics produced it is essential that
measures should be taken to ensure that the phase shift between lamp voltage
and transformer voltage is such that the harmonics produced in the trans-
former are in anti-phase with those produced in the lamp, so that they cancel
out; this is illustrated in Fig. 8.24, which only shows the third harmonics,
for the sake of simplicity.

Fig. 8.24. The zeroes of the lamp voltage and


transformer voltage should be shifted in phase so
that the higher harmonics cancel out.

If we display both the mains voltage and the lamp voltage on an oscillos-
cope, we see that the zeroes of the two curves are about 60 out of phase
when the total transformer voltage is 14 times the lamp voltage. This fact
restricts the application of this method, since a starter cannot be relied upon
to close when the mains voltage is only 14 x the lamp voltage. This is why
it is preferable to use two lamps in series, with an auxiliary capacitance
shunted across one of them. The starters then close in succession, each with
a voltage of 14 X (2 Yare) = 28 Yam which is generally quite enough for
reliable starting.
Example:
Ym = 118 V/50 Hz Two 40 W lamps in series
Dimensions of ballast 44 x 64 x 285 mm
Dissipation 23 W cos ff!tot = 093
Yrms in unloaded state: 284 V
8.4.4 Dual-lamp circuit
As we have just seen, the sequence start principle with a partial air gap
in the transformer core has advantages at low mains voltages (110-127 V)
when two lamps can be used in series. At high mains voltages (200-250 V),
142 FLUORESCENT LAMPS

on the other hand, where it is unnecessary to use a step-up transformer, it


is better to use alternate inductive and capacitive circuits; this has the addi-
tional advantage that the two lamps are independent of one another. Of
course, it is also possible to use a step-up transformer, with two lamps in
series; but the secondary winding is then generally so small that insufficient
harmonics are produced. It is then necessary either to take a lower tapping
so as to give an apparent increase in the size of the secondary winding, or
to introduce a separate self-inductance after all. These extra measures again
make the principle financially unattractive.
Various manufacturers have investigated whether coupling of the self-
inductances of the two lamp circuits can be advantageous. Fig. 8.25 shows
one quite commonly used version. The separate inductive and capacitive
sections can be clearly seen. Only a (small) part of the yokes is common
to the two sections.

__ .,.. __ , (l...!:-_-_-....-_-_----,
in d. cap.

Fig. 8.25. Coupling of a small part


of the iron circuits of the coils of
I I
the inductive and capacitive ballasts.

n
..-----....,-.. ; (-~-,1
I I I

'I b ' :

I
ind. cap.

Fig. 8.26. Economical coup-


ling of two iron circuits.

Fig. 8.26 shows a more elegant solution, in which the two coils share one
magnetic return path. Because the lamp current in the inductive circuit has
a phase lag of 60 (cos q; R:> f), while that in the capacitive circuit has a
phase lead of 60, the phase difference between the two lamp currents is 120.
LAMP TYPES AND CIRCUITS 143
The total current (mains current) of a dual-lamp circuit is thus equal in
magnitude to each of the separate currents. The same may be said of the
magnetic fields in the iron cores of the inductive and capacitive coils. This
means that we can always replace one or more sections of the separate iron
circuits by a common section of the same cross-section, as long as we take
steps to ensure that the two coils remain independent of one another.
Let us suppose that only lamp A is burning at a given moment. The
magnetic lines of force will then have the form indicated by the full lines a:
since the core of coil B contains an air gap, the magnetic resistance of this
part will be much greater than the two completely closed side arms, so that
very few lines of force will pass through coil B. Similarly, when only lamp B
is burning, the lines of force will have the form indicated by the broken
lines b. If both lamps are burning, the resultant field will be equal to either
of the individual fields of A or B. This use of a common return path thus
saves a good deal of iron, and also makes the losses less than with two
completely separated coils.
As we mentioned above, a small part of the magnetic field from coil A
does pass through B, and vice versa. This means that there is a definite (but
small) mutual inductance between the two branches of the dual-lamp circuit.
For example, if only the lamp with the phase lag (A) is burning, a small
voltage will be induced in coil B, thus making the open-circuit voltage across
lamp B somewhat higher; see Fig. 8.27, where 0-M represents the mains
voltage, and VB the voltage induced in coil Bas a result of the field in coil A.
The sum of these two voltages is applied across lamp B.

Vtamp 8
0 ..----------'------,

Votamp A Vcap.

Vchoke 8

Fig. 8.27 and 8.28. As a result of the magnetic coupling of


the coils of Fig. 8.26, the two lamp circuits are not
completely independent.

Conversely, it can be shown that when only the lamp with the phase lead
(B) is burning, the voltage across A will be slightly less (see Fig. 8.28).
The above-mentioned handicap can be completely eliminated by means
of a slight modification in the design of the circuit. For this purpose, the
144 FLUORESCENT LAMPS
v A
M
Fig. 8.29. The circuit with the phase
lead is here fed via a tapping on the
choke of the other circuit.

branch with the phase lead is not connected directly to the mains, but via
a small part of the coil with the phase lag (A); see Fig. 8.29, where the branch
with the phase lead is connected across 0-V, not across 0-M as usual. If
only the lamp with the phase lag is now burning (see the phase diagram of
Fig. 8.30), we find that the open circuit voltage across the lamp with the
phase lead, 0-V, is slightly decreased (distance 0-V smaller than 0-M).
Conversely, when only the lamp with the phase lead is burning (see Fig. 8.31),
the voltage across the other lamp rises somewhat (0-V is now greater than
0-M).

Vtamp B
0 ---~

I
I
I
Vcap.l
I
I
Ill.:
: 0 tampA
I
I
v I
I
I

Vchoke B
__ _t

Fig. 8.30 and 8.31. The arrangement of Fig. 8.29 causes


the two lamp circuits to influence one another in a man-
ner opposed to that caused by the magnetic coupling.

As a result of this special way of connecting the branch with the phase
lead, we thus get a certain interdependence between the two lamps. This
effect is opposed to that due to the coupling of the magnetic circuits; if the
tapping point V is correctly chosen, therefore, the two lamps can be made
completely independent of one another.
This gives a dual lamp circuit which works just like two separate circuits,
but which has only one iron return path, so that the dissipation is consider-
ably reduced.
Example:
Dual lamp circuit for two 40 W lamps Vmalns = 220 V/50 Hz
Dimensions 44 x 64 x 240 mm
Dissipation 14 W Im = 450 rnA cos cp;;:: 095
LAMP TYPES AND CIRCUITS 145
8.4.5 Ignition at various temperatures
Fig. 8.32 shows the mains voltage at which a 40 W lamp with starter circuit
will ignite, as a function of the ambient temperature. The increase of the
ignition voltage with decreasing temperature can be clearly seen; this is due
to a progressive reduction in the Penning effect (see Section 1.8.2).

\.{, (V)

1210
200

190

180

Fig. 8.32. Ignition voltage of 170


a 40 W lamp with an induc-
tive and with a capacitive
ballast, as a function of the -1 +10 20 30 40 50 60
ambient temperature. oc

8.4.6 Avoiding the luminous ripple


One advantage of the dual lamp arrangement is the pronounced reduction
in the luminous ripple arising from the fact that the emission of light from
the two lamps, like the lamp currents, shows a phase difference (see the oscil-
lograms of Fig. 8.33). The relative depth of the ripple in the combined light
is only half of that in the light from the individual lamps. Stroboscopic effects
are thus considerably reduced.
Luminous ripple can also be reduced to very small proportions by em-
ploying a three phase circuit: one lamp is connected to each phase of a
three phase system; the individual ripples of the three lamps then largely
counterbalance one another (see Fig. 8.34). However, this arrangement does
not give any correction of the power factor.

Fig. 8.33. Oscillograms of the


luminous flux of fluorescent lamps :
a. operated on choke ballast ;
b. operated on capacitive ballast ;
c. in dual-lamp circuit.
146 FLUORESCENT LAMPS

Fig. 8.34. Oscillograms of the luminous flux of I, 2 a nd 3 fluores-


cent lamps, each connected to one phase of a 3-phase system.

An economical 3 lamp circuit can be obtained by combining one branch


in which the current leads with two uncompensated inductive branches; the
power factor is then well above 08, with partial reduction in the luminous
ripple.

8.5. Lamp types for starterless circuits


Although the starter switch has proved to be a very useful component in a
fluorescent lamp circuit, there is a tendency to eliminate it. This would effect
a simplification from the maintenance point of view. In addition, the lapse
of time between switching-on and striking of the lamp can be considerably
reduced. These advantages have led to the development of a large variety
of starterless circuits.
For the majority of these circuits, described in the next section, new lamps
have been developed in order to achieve optimum results ; sometimes special
cathodes, capable of withstanding cold strarting, had to be designed. At the
same time there was a growing demand for lamps of higher light output,
and these lamps were for the greater part designed for starterless operation.
These trends in the development of fluorescent lighting have led to numerous
new lamps. Some of the most important types will now be discussed.
The most wide-spread types are those of the rapid start category. They have
been developed originally for operation on the rapid start circuit, with
cathodes suitable for 35 V preheat voltage. These lamps should be operated
with an earthed conductor (e.g. a fitting) at a distance not less than 2 em.
In the usual circuits reliable starting is not guaranteed at temperatures
below + 10 C.
Most types of starterless lamps require the use of some kind of starting
aid. Since the ignition kick supplied by the action of the starter switch is
not available, other means have been applied to facilitate striking. In many
cases this starting aid consists of a conducting strip or wire on the surface
of the lamps, extending from cap to cap. The metal shell of the lamp cap
LAMP TYPES AND CIRCUITS 147
should then be earthed. This is the system applied with the circuit of Fig.
8.16. A drawback of the necessity of earthing is that very often no reliable
earth connection is available.
The favourable effect of earthing the strip or a metal plate near the lamp
can be explained as follows. Since in most supply systems one of the con-
ductors is earthed, one of the lamp electrodes is also at earth potential. By
earthing the strip on the lamp a high potential difference is developed be-
tween the opposite electrode and the strip, which facilitates striking. In the
rare cases where the supply system is 'floating', i.e. without a connection to
earth, the strip on the lamp should be connected to one of the lamp electrodes
through a suitable resistor.
In the case of the circuit of Fig. 8.45 the corresponding lamp ('TL'M 40 W
RS) has an outside conducting strip which is connected to one of the elec-
trodes through a resistor of 1-2 MQ incorporated in the cap. No earthing
is required. The value of the resistor is high enough to avoid any danger
on touching the strip. In common with some other instant start lamps, the
'TL'M lamp has a silicone coating which makes its striking properties inde-
pendent of the humidity of the lamp surface or the surrounding atmosphere.

240 285

I K-~
Vo

I
Vm
260
/
v
200 236

212
/
160
""' I'-
--- 188

165

120 142
Fig. 8.35. Variation of the strik-
ing voltage with ambient tem-
perature for a 40 W 'TL'M lamp -20 0 20 40 60C
on a 220 V 'TL'M ballast. - - - Ambient temperature

The 'TL'M lamp, on its corresponding ballast, ignites reliably at low tem-
peratures, which makes this combination particularly suitable for outdoor
applications. Fig. 8.35 gives the relationship between striking voltage and
the ambient temperature for a 40 W lamp. The ballast commonly used with
this lamp is of the semi-resonance type (see Fig. 8.45). The open circuit
voltage, indicated on the right hand side of the graph, is higher than the
mains voltage, indicated on the left hand side. The curve shows a peculiarity
which is common to all fluorescent lamps, viz. the fact that also at higher
temperatures striking becomes more difficult, so much so, that for bulb wall
temperatures above 60 C more than rated voltage is required.
Another special lamp for starterless operation is the Philips 'TL'S lamp*.

* W. ELENBAAS and T. HoLMES, Philips Techn. Rev. 12, 129 (1965).


148 FLUORESCENT LAMPS

This lamp is provided with an internal conducting strip, one end of which
is connected to an electrode. When voltage is applied, a glow discharge
occurs in the lamp between the free end of this ignition strip and the other
electrode and this produces enough ions to start the main discharge between
the electrodes themselves. Very little current flows along the path provided
by the ignition strip, whose resistance is fairly high, and the 'strip losses'
are accordingly small. Together with the optical losses caused by the strip
they amount to approx. 15% of the total consumption of the lamp. The
electrical losses occur when the strip is positive with regard to the discharge.
This means that unequal amounts of light are radiated in the two halves of
a cycle. The light thus contains a 50 Hz ripple which is superimposed on the
normal 100 Hz ripple and may occasionally be observed by the eye.
In two-lamp fittings it is for this reason advisable to connect the wiring
in such a way that the 50 Hz components in the light of both lamps counter-
balance.
Lamps of the inside strip type can be operated in series with a choke
ballast or a resistor which may also take the form of a tungsten lamp.
Ignition is quicker in the latter instance, so much so in fact that the lamp
may be considered as lighting instantaneously. The 'TL'S 40 W lamp is
marketed for general lighting purposes and, in conjunction with a tungsten
lamp, it ignites promptly on 200 V a.c. If a choke ballast is employed,
ignition is reliable only on mains supplies when the voltage does not drop
under 220 V.
The fact that the internal ignition strip is connected to one of the electrodes
makes the use of a special lamp cap preferable, since, in the event of only
one end of the lamp being inserted in the holder and the contacts at the
other end being touched, the person concerned could be exposed to the full
force of the mains voltage, because the circuit is completed through the gap
between the strip and the unconnected electrode.
A specially designed lamp cap (see Fig. 5.9c) rules out this possibility
altogether.
The dimensions of the 'TL'S lamp are such that the overall length of the
mounted tube with holders is the same as that of the standard 'TL' 40 W
lamp with its appropriate holders.
In the 'TL'S lampholder the contact is capable of being rotated and, in
the position shown in the photograph, it carries no current; the circuit is
completed only when the lamp is inserted and rotated in the holders. See
Fig. 8.36.
'TL'S 40 W lamps give a luminous flux of approximately 2 500 lm and
the light is of the white or the warm white colour which blends well with the
light emitted by the tungsten-filament lamp burning in series with it.
Apart from the advantages of immediate ignition, the omission of the
starter switch is a good thing in that it makes possible the use of fluorescent
lamps in locations where explosion hazards exist, such as in chemical works,
coal mines and so on, where the use of the standard 'TL' lamps with starter
might otherwise be dangerous. For, if a lamp is broken, but the electrodes
remain intact, the starter will render the electrodes incandescent and might
thus cause an explosion. For use in coal mines the 20 W 'TL'S lamp has
been developed and, as there is usually some freedom of choice in the matter
LAMP TYPES AND CIRCUITS 149

Fig. 8.36. Lamp cap and holder of the 'TL'S lamp.

of supply voltage, this can be made sufficiently high (200-220 V) to guarantee


immediate ignition. Further, these lamps will re-ignite at once after being
switched off, this being important when the lamps are also used for signalling
purposes.
In order to take full advantage of the use of 'TL'S lamps in explosive
atmospheres a special single pin cap and corresponding holder have been
developed. In this execution the lamps are designated as 'TL'X lamps (see
Fig. 5.9d). The idea of the cap and holder design is to prevent explosions
when contact is interrupted in the holder. This happens in a small explosion
proof chamber; the pin fits closely in an entrance opening, thus preventing
a small explosion in the chamber from igniting the surrounding atmosphere.
Another method of reducing the ignition voltage of a lamp, is the applica-
tion of a transparent conductive layer on the inside of the tube wall. Many
technological difficulties had to be solved, mainly related to the attack of the
layer by mercury atoms, but the ultimate result is a lamp with excellent igni-
tion properties and reasonably good maintenance characteristics.
In Fig. 8.37 the ignition voltage as a function of the ambient temperature
is given for a few typical lamp types.
The IEC has standardised a circuit for measuring the ignition voltage of
lamps with preheated cathode. It is given in Fig. 8.38.
The cathodes of the lamp are connected to a system of variable trans-
formers which permits the adjustment of the required preheat voltage.
If specified, an earthed starting aid (e.g. a metal plate) can be applied.
The voltmeter Vs indicates the voltage obtained from the variable supply,
which should be regulated in such a way that the minimum voltage at which
150 FLUORESCENT LAMPS

V0 260

t 220

180

160

140
L--+~~~~-+~~~~~~
-15 -5 0+5 15 25 35 45 55
-oc Fig. 8.37
the lamp starts, is delivered.
'Slimline' lamps are manufactured in various type sizes. At first, lamps of
a small diameter were made, for which the name 'Slimline' was coined. The

Table 8.4 Data of Slimline lamps

lamp size approx. luminous flux


watts arc voltage current (Cool White)
48x1tin 38 97 425 rnA 2 900 lm
72X 1 in 37 210 200mA 2 950 lm
72x1tin 55 145 425 rnA 4 400 1m
96x 1 in 49 285 200mA 4100 lm
96x It in 74 192 425 rnA 6 000 1m

Fig. 8.38. Circuit for checking the starting voltage of fluorescent lamps.
LAMP TYPES AND CIRCUITS 151
demand for larger lighting units led to bigger lamps, with the result that at
present most 'slimline' lamps are made in the normal 11- in diameter.
Table 8.4 gives a survey of some of the most important types, together
with their electrical characteristics and light output.
These lamps are provided with a single pin contact. The corresponding
lampholders are designed in such a way that the circuit is completed by
inserting the lamps. Without the inserted lamp live parts in the holders can
not be touched. No special starting aid is provided for these lamps. Reliable
striking is ensured by the use of a sufficiently high no-load voltage of the
ballast.
The rapid start lamps, referred to in the beginning of this section have
the same lumen output as the normal fluorescent lamp of the same wattage.
They are also provided with the normal hi-pin cap. The range of rapid start
lamps has been extended with a number of high output lamps. These are
equipped with a special cap with recessed contacts. The principal data of
these lamps are shown in Table 8.5.
Table 8.5 Data of high output lamps
lamp size watts arc voltage
I current

48 X It in 60 80 800mA
72x It in 85 115 800mA
96x It in 105 148 800 rnA

The demand for lamps of still higher lumen output led to the development
of the very high output lamps. In developing these lamps, special measures
had to be taken for keeping the mercury vapour pressure within reasonable
limits. Various solutions led to the required results as indicated in Section
2.5. These lamps are designed to operate at 15 A. Table 8.6 provides the rele-
vant data for some of the most important types.
Table 8.6 Data of very high output lamps

lamp size arc voltage lamp current approx. luminous


type
flux (Cool White)

'TL'M 115 W/RS 60xltin 95 15 A 7100 lm


'TL'M 140 W/RS nx It in 114 15 A 8 500 lm
'TL'M 215 W/RS 96x It in 170 15 A 13 000 lm

In the above lamps the mercury pressure is controlled by a cool space


behind the cathodes, as indicated in Fig. 2.4. Another series of lamps uses
tubes with a non-circular cross-section as indicated in Fig. 2.3 (Power-
Groove type). These lamps cover roughly the same range as those men-
tioned above.
152 FLUORESCENT LAMPS

8.6 Circuits for preheat starterless lamps


It will be clear that the ballast in these circuits has the same stabilising
function as in the circuits described in Section 8.4. All the points mentioned
there must be taken into consideration here too, e.g. dissipation, steady
burning despite variations in the mains voltage, power factor, etc. It follows
that all starterless circuits can basically be considered as derived from some
type of circuit with starter, with either inductive or capacitive stabilisation;
and all the considerations and characteristics which apply to the one will,
mutatis mutandis, apply to the other.
These starterless circuits have the additional function of providing the
appropriate ignition conditions, which generally comes down to: (a) applying
the correct ignition voltage, and (b) heating the electrodes to emission tem-
perature, by means of the appropriate heater current.

8.6.1 Relation between open-circuit voltage and cathode current


If a given current Ik is passed through the cathode of a fluorescent lamp
until the equilibrium temperature is reached, the voltage V0 at which the
lamp ignites can be determined. The relation Vo = f(ik) is sketched in
Fig. 8.39; see also the text relating to Fig. 1.12. To the left of ie, the cathode
is too cold to emit electrons; under these conditions, a high voltage V1 is
needed to ignite the lamp.

Fig. 8.39. Schematic representation ofthe


ignition voltage as a function of the catho-
de heating current.

With ik ~ ie, the emission temperature is attained or exceeded, and the


ignition voltage is lower. The voltage difference V1 - V2 is clearly the
potential difference needed to extract electrons from the cold cathode. The
lamp will thus ignite reliably with a voltage above V1o or above V 2 if the
cathode current ik ~ ie. However, it is not advisable to ignite the lamp with
too cold a cathode: a high voltage ~ V1 not only pulls electrons out of
the cathode, but also damages the emitter layer on this electrode. Small
particles of the emitter are dislodged, and are deposited on the wall of
the lamp, where they show up as dark spots. This process limits the life of
the cathode, and hence of the lamp.
LAMP TYPES AND CIRCUITS 153
It is thus advisable to choose the open circuit voltage between V1 and V2 ,
so that the lamp only needs to warm up for a short time. Lamps using this
system are called 'rapid start' lamps, as opposed to 'instant start' lamps,
which use voltages greater than V1
It should be realised, however,' that the curve of Fig. 8.39 holds only for
one ambient temperature. At a lower ambient temperature, the curve as a
whole will shift to higher voltages. If the lamp is to be designed for use at
low temperatures, the appropriate V0 - ik curve must be taken into con-
sideration. However, this can lead to conflicting demands on the lamp:
as the V0 - ik curve rises with decreasing ambient temperature, a moment
will come where V2 low temp > V1 normal temp If now the lamp is given such
a high open-circuit voltage that it ignites reliabily at low temperatures this
entails the great disadvantage that the lamp will ignite with a cold cathode
at normal temperatures, and will thus have a shorter life. If the ignition
voltage is lowered, ignition will be poor at extremely low temperatures.
The relation between the minimum ignition voltage (assuming adequate pre-
heating) and the ambient temperature, V2 = f(T), is given in Fig. 8.35.

8.6.2 Series and parallel heating of the electrodes


The cathodes can be heated in two different ways:
( 1) In parallel heating, the cathode is connected directly to a voltage source
of low internal resistance; see e.g. Fig. 8.42. In this case, the current is
inversely proportional to the resistance of the tungsten filament, which
is appreciably lower when the filament is cold than when it is hot. The
filament heats up quickly under these conditions.
(2) In series heating, the filaments are heated in series with a large impe-
dance. This makes the current constant (independent of the resistance
of the filament); but the cathode heats up more slowly, so that the lamp
also ignites more slowly. An example of this type may be seen in Fig.
8.46, where the filaments are pre-heated in series with a capacitance.
While series heating thus has the disadvantage of giving slower ignition,
it has the advantage of more reliable operation, because:
(1) contact resistances play hardly any role;
(2) If the current is interrupted for any reason, both cathodes are without
current and the lamp does not ignite at all. In parallel heating, on the
other hand, if one of the cathodes is not heated or is insufficiently heated,
the lamp will still ignite, but initially it will burn asymmetrically (current
only flows in one direction); in the other half-period, the lamp will ignite
cold after a few seconds, with the above-mentioned disadvantage of a
shorter life.
(3) With parallel heating, a short circuit of the leads near the lampholder
can lead to overloading of the heating transformer which may cause
irreversible damage.
However, the choice of the kind of heating to be used depends even more
on the costs involved than on the above-mentioned factors. In some cases,
especially when the mains voltage is low and long lamps are used, parallel
154 FLUORESCENT LAMPS

heating can be realised quite cheaply; in other cases, especially with mains
voltages of 220--250 V, series heating is the logical solution.

8.6.3 Influence of the choice of the cathode on the ballast


Years ago, nearly all fluorescent lamps had 'high resistance' electrodes,
which gave a voltage drop of about 8 V when the cathode current delivered
by the ballast passed through them. Since the advent of starterless circuits,
however, the need for a low resistance filament, with a voltage drop of about
3-! V, has made itself felt. The main reasons for this are as follows:

(1) The voltage applied across the filaments is maintained, at least in part,
during the burning of the lamp, unlike the case e.g. with glow switch
starters, where the starter is non-operative during the burning of the
lamp. This continued heating leads to unnecessary dissipation, so it was
natural that attempts should be made to reduce the ohmic resistance of
the heater filaments.
(2) If the nominal electrode voltage is 8-9 V, mains fluctuations can cause
the voltage across the filament to exceed this value. However, the catho-
des in the old lamps were so short that a transverse discharge was pro-
duced between the filament leads (Fig. 8.40) under these conditions. This
(gas) discharge had to be stabilised, just like the main discharge through
the lamp, which complicated the cathode heating system.
The risk of such a transverse discharge is avoided if the filament
voltage required can be made sufficiently low.

Fig. 8.40. If the voltage across the electrode ex-


ceeds 10 V, a transverse discharge can be produced.

(3) In starterless circuits, the open circuit voltage is applied at the same
time, as the heating current, unlike the case with starter circuits, where
the high ignition-voltage peak does not occur until the starter switch
opens, at which time the electrodes are already pre-heated. If the open
circuit voltage is chosen fairly high, so that the lamp can operate in a
wide ignition range, the total number of times a lamp can be switched
on during its life is less than with a bimetallic starter. A triple coiled
filament, on which more emitter powder can be applied, has solved this
problem.

One advantage of the starterless circuit is that the ignition of the lamp is
made more reproducible. This leads to a better-looking life curve (Fig. 8.41 ),
which means that group replacement of lamps is a more attractive proposi-
tion. With starter circuits, the life of the lamps is more variable, because of
LAMP TYPES AND CIRCUITS 155
Number
of lamps
%
f +.,~00~----------~,
Fig. 8.41. Number of lamps still operating,
as a function of the operating time. Curve I
applies to lamps with starterless circuits, and 50
curve II to starter lamps. The fact that the
ignition process is reproducible with starter-
less circuits reduces the spread in the break-
down rate.
-time

the greater spread in starter properties (speed of opening, etc.). This can be
a very real advantage in large installations, where the more efficient group
replacement leads to appreciable savings.
We shall now discuss the various stabilisation methods mentioned m
Section 8.4, to see which of them are applicable in the present case.

8.6.4 L and L-C circuits with cathode transformer


If a single self-inductance is enough to stabilise the discharge, It IS an
obvious idea to add a transformer to provide the heating current. This
system has indeed proved to work quite well.
The primary of this auxiliary transformer Tr can be connected to the mains
terminals, but it is better to connect it in parallel with the lamp as shown
in Fig. 8.42; this will lower the transformer voltage to the lamp voltage
during operation. This leads to a substantial reduction in the (superfluous)
heating currents supplied to the electrodes during operation.
A pre-heat transformer can also be used with a capacitive ballast, but only
across the mains (Fig. 8.43): if the transformer is connected across the lamp,
it can overheat so much as to cause permanent damage at the end of the

Fig. 8.42. Rapid-start circuit using a start- Fig. 8.43. Capacitive circuit for rapid-start
ing transformer for preheating the electro- lamps.
des. A small capacitor is shunted across the
lamp to suppress radio interference.
156 FLUORESCENT LAMPS

lamp's life. This is due to the fact that at the end of a lamp's useful life the
lamp is in effect a rectifier, since one of the electrodes has lost its emissive
power. The resulting asymmetrical lamp current has a d.c. component which
cannot pass through the capacitor; it therefore all flows through the primary
winding of the pre-heat transformer, leading to serious overloading.
In an inductive circuit, nearly all the direct current will flow through the
choke, which has a much lower d.c. resistance than the transformer.
Two pre-heat transformers can be used for a dual-lamp circuit (with one
inductive and one capacitive circuit). It is however possible to economise
by using a single transformer as shown in Fig. 8.44. This transformer is
connected across the mains, so that there is no drop in the heater currents
after the lamps have started.

Fig. 8.44. Dual-lamp circuit for rapid-start


lamps, using one preheat transformer.

Applications

(a) 'TL'M or 'TL' /RS lamps with low-resistance electrodes

The mains voltage for these lamps must be so much more than the lamp
voltage that an inductive or capacitive circuit can be used without a step-up
transformer after the lamp has ignited. In practice, this condition is generally
satisfied by mains of about 220 V if the lamp rating does not exceed 40 W;
however, this circuit does often suffer from unreliable ignition.
The temperature ranges within which 40 W lamps will ignite reliably on
the 220 V mains (assuming a negative mains fluctuations of 10%) are:
-10 oc < 'TL'M 40 W < +35 oc
+17 oc < 'TL' 40 W/RS <+52 oc
Neither of these ranges is really enough to cover many cases met with in
practice. With 65 W lamps, the temperature range within which ignition is
reliable is so narrow that this solution is not a practical proposition.
LAMP TYPES AND CIRCUITS 157
(b) Lamps with high-resistance electrodes
These lamps are found especially in British Commonwealth countries at
present. In areas with mains voltages of 240 V, a circuit like that shown in
Fig. 8.42 is enjoying some popularity because of the relatively wide tempera-
ture range in which the lamp will strike. However, the combination of a
capacitive ballast and a pre-heat transformer is not to be recommended with
these lamps.
The rather high electrode voltage required may, as mentioned above, give
rise to a gas discharge across one or both electrodes, and the discharge cur-
rent must be limited to a safe value by a stabilising impedance. In Fig. 8.42,
this function is performed by the choke L; In Figs. 8.43 and 8.44, however,
the transformer is directly connected to the mains without any stabilising
impedance in series. The resulting discharge current can be very damaging
to the electrodes. For this reason, inductive stabilisation is nearly always
used with high resistance electrode lamps. A capacitor is then shunted across
the mains to give a high power factor.

8.6.5 Semi-resonance circuit


The temperature range in which starterless lamps strike when the open
circuit voltage is equal to 220 or 240 V (the mains voltage in many countries)
is often too narrow for practical use. For this reason, Philips have developed
a circuit on an entirely new principle for these conditions.
The operation of this circuit can be explained starting from the compen-
sated inductive circuit shown in Fig. 8.45a together with its vector diagram.
The total mains current for this circuit is equal to the vector sum of the
lamp current 11 and the current through the compensation capacitor, lc. If
the iron core of the choke is now given a second winding with an equal
number of turns, which will be tightly magnetically coupled to the choke
coil, we get the connection for the compensation capacitor shown in Fig.
8.45b, which will have precisely the same vector diagram as that of Fig.
8.45a: since the number of turns of S 1 = number of turns of S2 , and since
these two coils are connected at one end (via the cathode), there can be no
voltage difference between the other ends of the coils (points A and B). Since
the potential of A is always equal to that of B, the capacitor can be con-
nected to B without changing lc.
The mains current now flows through S 1 , and the capacitor current
through s2.
Since Im and Ic combine to give the lamp current (Fig. 8.45a), the magnetic
field in the iron core due to the ampere-turns of sl and s2 will be equal to
that due to the ampere-turns of S in Fig. 8.45a (the lamp current also flows
through S). We may thus write Vs = Vs 1 = Vs 2
The advantages of the circuit of Fig. 8.45b manifest themselves in the
unloaded state. Current then flows through S 1 and S 2 in opposite directions,
158 FLUORESCENT LAMPS

5 s,

.... ....

.9. b ..

.:siH
vm

L
/1)

i"c Im
I c'
/rm
It
vt
Fig. 8.45. Modification of the normal compensated circuit (a) to give a semi-resonance
circuit (c). Circuit (b) is an intermediate stage.
so that the magnetic fields of these two coils cancel out. A current which is
determined practically entirely by the impedance of the capacitor flows
through the electrodes, which can thus be pre-heated. Although the circuit
of Fig. 8.45b thus represents a suitable means of satisfying the two ignition
conditions (ignition voltage across the lamp and preheating current through
the electrodes), it needs to be modified slightly to give optimum performance.
Since the influence of S 1 is equal and opposite to that of S 2 , the available
open circuit voltage will be no greater than the mains voltage, and since fc
is always less than / 1 (see the vector diagram of Fig. 8.45a), the heater cur-
rent will also be quite small. A particularly elegant solution for this problem
is shown in Fig. 8.45c; if the coupling between S 1 and S 2 is made less than
100% (e.g. by winding the two coils next to one another, as for the leak
transformer shown in Fig. 8.12, Section 8.4), an extra (leak) reactance will
be produced in series with both S 1 and S 2 . These extra reactances do not
contribute to the transformer effect, and can thus be represented by separate
impedances SIH and s2H
This circuit has two advantages in the unloaded state. Firstly, the heater
current is now determined by the differences of the impedances Zc-
(Zs 1 H + Zs 2H), and is hence larger. Further, the open circuit voltage is no
longer equal to the mains voltage vm, but to vm + fo. ZstH By suitable
choice of these two impedances, therefore, both the open circuit voltage and
the heater current for the electrodes can be given the values that suit the
lamp best.
The vector diagram under operating conditions does not differ too much
from that shown in Fig. 8.45a; we merely have to insert the voltage vectors
Vs 1 H (perpendicular to Im) and Vs 2 H (perpendicular to fc).
LAMP TYPES AND CIRCUITS 159
Thanks to these two extra 'coils', this circuit also has two other surprising
advantages, which can be clearly seen from the numerical example of Fig.
8.46. This circuit applies to a 40 W 'TL'M lamp and 220 V/50 Hz mains.
(1) The higher harmonics generated in the lamp current (see also Chapter?)
are strongly attenuated in the mains current, because the two extra coils
and the capacitor combine to form a low pass filter. The current drawn
from the mains is thus more or less sinusoidal, which is an unusual
property for a circuit with a high power factor.
(2) The overall impedance of the circuit for audio frequencies, such as those
used for signalling and measurement purposes in the mains, is high.

BOnil 400nil
BOil

BOil 4001l

Fig. 8.46. Values of components in the semi-resonance circuit for a 40 W lamp on the
220 V and 50 Hz mains.

Applications
This circuit can be used with mains voltages from 200 to 240 V and lamps
up to 120 W, i.e. in cases where normal inductive or capacitive stabilisation
would be used in starter circuits.
Example:
One 40 W/RS 'TL'M lamp Vm = 220 V/50 Hz
Max. dimensions of ballast: 44 X 64 X 240 mm
Dissipation 10! W cos q; = 093 Im = 240 rnA
Cathode heating current 450 rnA
Open circuit voltage 260 V
Temperature range for reliable ignition at nominal mains voltage: -20 ac
to +60 C.
160 FLUORESCENT LAMPS

Fig. 8.47. Modifications of the semi-reso-


nance circuit.

8.6.6 Modifications of the semi-resonance circuit


When designing semi-resonance circuits of the type described above, we
must take many different factors into consideration if optimum striking and
burning of the lamp are to be achieved. It is not always possible to satisfy
all demands simultaneously.
For example, if the ballast has to be made thinner than usual, difficulties
will be encountered making the leak impedances right; these will tend to be
either too big or too small, unless special measures (involving additional
expenses) such as splitting each winding into still more sections are taken.
In such cases, good results can be obtained with the circuits of Fig. 8.47.
The capacitor current here flows through the lower electrode, but a special
current transformer is included in the circuit to supply the heater current
for the upper electrode. The coil Scan be a normal choke, as used in starter
circuits. If desired, the open circuit voltage and the heater current can be
increased by adding the auxiliary capacitance Cn to the circuit. The series
combination of S and Cn then serves to give a higher voltage across Cn,
and hence across the lamp. When the lamp has started, the voltage across
Cn falls to the lamp voltage, as a result of which the heater current also
decreases.
Example: (see Fig. 8.47b)
One 40 W 'TL'M lamp Vm = 220 V/50 Hz
Dimensions of ballast 38 x47 x278 mm
121- W losses
Open circuit voltage: 250 V
Cathode heating current 430 rnA
Another difficulty arises if the ratio Vm/V1 falls from the preferred value
of about 2 to 15 (this problem has already been discussed in Section 8.4).
A circuit with one lamp and a single choke will then not be stable enough:
slight variations in the mains voltage will give rise to excessive variations in
LAMP TYPES AND CIRCUITS 161

Fig. 8.48. Modification of the semi-resonance circuit. The stabilisation coil is also used
as a transformer to increase the open circuit voltage. In fig. (a), series heating is used for
both electrodes, while in fig. (b) parallel heating is used for the upper electrode.

the lamp current. A circuit like that shown in Fig. 8.48a or 8.48b may be
the solution in this case. These two circuits differ in the means used to heat
the upper electrode; they both have the same vector diagram (Fig. 8.49).
The main difference between this and the semi-resonance circuit described
above is that here C is connected to a tapping some way along S. The com-
bination of Sa and C (and possibly Lh) gives an increase in the voltage across
this capacitor. The transformer effect of .the windings Sa and Sb increases
the open circuit voltage across the lamp to the desired value. When the lamp
is burning, it is connected in series with S. Now the magnetic field of S is
determined not only by the ampere-turns Sab . ! 1 but also by Sb . Ic. The
resultant field is oriented so that the voltage vector, at right angles to it, is
no longer at right angles to the lamp current; this is illustrated in Fig. 8.49.
The sum of V 1 and Vs is equal to V It may be clearly seen that the lamp
111

apparently burns at the open circuit voltage found in the unloaded state, V0
This voltage is appreciably higher than the mains voltage, so that the lamp
current will be sufficiently stable.
Seen from the lamp, the capacitor is in parallel with part of the self-
inductanceS, which means that the impedance for higher frequencies (higher
harmonics) will be too low. Both the lamp current and the mains current

Fig. 8.49. Vector diagram of the circuit of Fig. 8.48.


162 FLUORESCENT LAMPS

can be excessively deformed under these conditions. This is the reason for
the inclusion of Lh in the circuit. This auxiliary coil can if desired be placed
in the mains lead, and/or used as a voltage source for heating one or both
of the electrodes.
Example (see Fig. 8.48b):
One 105 W 'TL'/RS lamp Vm = 220 V/50 Hz cos cp = 097
Dimensions of ballast 44 X 64 x 375 mm
Watt losses: 28 W V1 = 150 W / 1 = 875 rnA
Unloaded: V0 = 360 V / 0 = 1100 rnA

8.6.7 Auto leakage transformers


The taking of tappings from stabilisation coils cannot be taken too far.
As Vm/V1 falls, the sensitivity to mains fluctuations increases sharply. The
dimensions of Lh also increase sharply, and the circuit soon becomes un-
economical.
In the very important range of application where Vm R:> V 1 (think e.g. of
40--80 W lamps on 110-127 V mains), the auto leakage transformer is the
best solution. The change from starter circuits to starterless circuits makes
least difference in this connection. If we use the existing auto leakage trans-
former for the corresponding starter circuit, all we have to do to convert it
to a starterless circuit is (see Fig. 8.50):
1. to provide two extra tappings for pre-heating the cathodes,
2. to increase the open circuit voltage somewhat, depending on the desired
ignition range.
The price difference between starterless and starter circuits is so slight
here that starter circuits are now hardly ever used if the circuit has to include
an auto leakage transformer anyway.

Fig. 8.50. Auto leakage transformer used Fig. 8.51. Single-lamp rapid-start circuit
in a rapid-start circuit. with high power factor.
LAMP TYPES AND CIRCUITS 163
Example (see Fig. 8.50):
One 40 W 'TL'/RS lamp Vm = 118 V/50 Hz
Watt losses: 14 W
Max. dimensions of ballast 38 x47 x 150 mm cos fPtot = 045
Open-circuit voltage: 232 V
Cathode voltage (unloaded): 37 V

While the use of an auto leakage transformer as such is very simple, the
costs of compensating the circuit so as to give a high power factor are con-
siderable. As we have seen above, capacitors, shunted directly across the
mains terminals give a lot of problems, while capacitors for low voltages
cost relatively more than those for high voltages and the same power. This
last problem may be solved by making use of the circuit of Fig. 8.51 where
the capacitor is connected across a higher voltage so that it can be made
smaller; on the other hand, this involves the application of an extra step-up
winding. If a dual lamp circuit is required anyway, it may be more economical
to make use of the sequence start principle (which automatically gives a high
power factor).

8.6.8 Sequence start circuit


Here again, the simplest way to derive the starterless circuit from the
sequence start circuit with starters (whose theory has been discussed
Section 8.4) is to replace the starters by three windings for pre-heating the
electrodes to emission temperature (compare Fig. 8.19 and 8.20 of Section
8.4 with Fig. 8.52). The lower lamp ignites in series with C 1 , and draws a
current of only a few milliamperes. The voltage drop across C (C C1 )
is now so small that practically the whole open circuit voltage is applied
across C1o so that the upper lamp ignites too. As with the corresponding
starter circuit, it is desirable that the open-circuit voltage should be about
14 times the total arc voltage, to ensure steady burning. This voltage is
generally on the high side for the ignition of the first lamp, and can shorten

I
I
I

c2:;:
I

I
I
Fig. 8.52. Popular starterless circuit working I
I
on the sequence-start principle. The distortion
of the lamp current is balanced by the counter-
distortion produced in the transformer.
164 FLUORESCENT LAMPS

the life of this lamp. The voltage divider C 1 - C 2 improves the situation
by adjusting the open circuit voltage across the first lamp to the right value,
while also serving to suppress radio interference.
Since most installations have more than one lamp per fitting in any case,
the circuit of Fig. 8.52 is the solution chosen in those countries where a
transformer has to be used anyway.
However, this circuit is not suitable for use at low temperatures, as it
contains no self-inductance in series with the capacitor to suppress higher
harmonics, apart from the 'spread' self-inductance caused by the juxta-
position of the primary and secondary windings. As a result of this lack of
a self-inductance, most sequence start circuits in common use give difficulty
with starting at temperatures below about 5 C.
Example (see also Fig. 8.52):
Two 40 W 'TL'M lamps Vm = 118 V/50 Hz
Dimensions of ballast: 44 x 64 X 285 mm
Watt losses: 22! W cos rp = 095
Open circuit voltage: 285 Yrms Cathode voltage: 37 V
Ignition range at 90% nominal mains voltage: +3 octo +47 C.
Ignition range at nominal mains voltage: -20 oc to +57 oc (but tends
flicker at low temperatures)

8.6.9 Ignition with the aid of semiconductors


The advent of reliable semiconductor devices provides another possible
way of igniting fluorescent lamps by means of an auxiliary circuit which is
then cut off.
One field in which the use of such circuits is attractive is that of the longer
(8 ft) lamps with arc voltages from 160 to 180 V. Once these lamps have
been started, they can quite simply be stabilised with a simple L-C series
circuit with mains voltages ~ 220 V; but before the use of semiconductors
these lamps had to be provided with transformers or resonance circuits for
ignition purposes, and these continued to supply the lamp current (at least
in part) even after ignition. These ballasts are so large and expensive, and
have such a high dissipation, that they have never been very popular. They
can now be replaced by an electronic system which can be made much
smaller and has lower dissipation.
The principle of this method is illustrated in Fig. 8.53. L 1 represents the
impedance for stabilising the lamp current; this can be made an L-C com-
bination if desired, without affecting the principle of operation of the circuit.
The starter proper consists of three parts, the capacitor C2 , the self-induct-

Fig. 8.53. Use of an electronic ignitor gives a simple


stabilisation circuit (C1 - L 1 ), thus reducing the
dimensions and dissipation of the ballast.
LAMP TYPES AND CIRCUITS 165
ance L 2 and the switch S. When Sis open, C2 and L 1 form a tuned circuit,
whose resonance frequency should be made high compared with the mains
frequency. This circuit is designed to charge C2 periodically to the ignition
voltage of the lamp. When Sis closed, on the other hand, C 2 forms another
tuned circuit, with L 2 ; this has an even higher resonance frequency (L 2 L 1 ),
so that the effect of L 1 on this circuit is negligible. This circuit causes C2
to be discharged, so that the electrodes of the lamp are heated to emission
temperature.

-1

Fig. 8.54. The electronic ignitor con-


tains two thyristors in anti-parallel,
each with its triggering circuit.

In fact, the 'switch' S consists of two thyristors connected in anti-parallel


(Fig. 8.54). These are triggered in a normal way by an auxiliary capacitor
(e.g. C3 for T 3 ), which is charged via the voltage divider R 1 -RrR 3 After
the breakdown voltage of the diac D has been reached, the voltage across
the diac falls sharply, and the capacitor in question is partially discharged
across the gate-cathode junction of the thyristor, triggering the latter. If we
assume for the sake of simplicity that the circuit is switched on at time t0
(see Fig. 8.55), then the voltage across C2 will remain practically equal to
the supply voltage (____!__C w L 1 ). This voltage is also applied across the
w 2
voltage divider R 1 -RrR 3 , and is used to charge the two triggering capaci-
tors. The time constants are chosen so that one of the thyristors (e.g. T3 )
is triggered at time t 1 . This closes the circuit C2 -L 2 , which oscillates with
high frequency (e.g. 1 000 Hz) and decreasing amplitude, due to the dam-
ping of the electrodes of the fluorescent lamp, which are heated by this means.
The variation of current and voltage with time is shown in Fig. 8.55.
When the current passes through zero (half a cycle later.), the thyristor is
cut off again, and we now get a completely new situation, in which C2 (with

-----...........
:
''
Fig. 8.55. Variation of the capa- '
'
citor voltage Vc 2 with time during
to
' I

t3 i4
I I I 1

the ignition of the lamp. t( t2


166 FLUORESCENT LAMPS

a negative voltage Vc2t 2 ) is connected via L 1 to the mains, which has an


instantaneous value of Vmt 2 This causes C 2 to be charged with a frequency
which is high compared with the mains frequency (e.g. 400Hz) to a voltage
Vc 2 t 3 such that Vcu 3 - Vmt 2 = Vmtz- Cc 2 t 2 (assuming that the losses in
the circuit are negligible, and that the mains voltage has not changed in
the short time interval between t 2 and t 3 ).
By time t 4 , the auxiliary capacitor C3 is again charged to the breakdown
value of the diac, so that the thyristor is triggered again. The tuned circuit
C 2 -L 2 is then closed, and the whole process is repeated as after time t 1
It may be remarked that the first charging time of the auxiliary capacitance C3
(from t0 to t 1 ) is much longer than the second and subsequent charging times.
This is understandable, as C 3 has to be charged right from zero to the break-
down voltage of the diac the first time, while after the thyristor has been
triggered the capacitor still has a certain residual voltage ( = forward voltage
of diac).
This process of resonance charging of C2 via L 1 and discharging via L 2
continues a number of times, until a moment comes when C 3 is no longer
charged to the breakdown voltage of the diac. This occurs near the end of
each half cycle of the mains voltage, when the instantaneous value of Vm
is no longer sufficient to charge c2 to the required value.
The open circuit voltage across the lamp is thus formed by the mains
voltage with voltage pulses of higher frequency superimposed on it. After
the lamp has ignited, the arc voltage has more or less the form of a square-
wave, and if the voltage divider RrR 2 -R 3 is suitably designed the voltage
across C 1 and C 3 will remain lower than the breakdown voltage of the diacs.
Under these circumstances, the thyristors will remain cut off, so that the
starter 'switch' will be out of operation.
Example:
One 85 W 'TL'/RS lamp Vm = 220 V/50 Hz
V 1 = 178 V / 1 = 560 rnA
Watt losses: 12 V cos cp = 080 capacitive
Unloaded: Eo peak = 700 V Ieath = 600 rnA
Dimensions of ballast: 44 x 64 x 230 mm

8.6.10 Circuits for extremely low temperatures


Most ballasts provide reliable ignition down to a temperature of 0 C,
while in a few cases (e.g. with the semi-resonance circuit) the ignition range
extends down to -20 ac; this is needed for street lighting. However, in cer-
tain special applications (e.g. deep freeze chambers) ambient temperatures
of -40 ac - -50 ac are quite usual.
The variation of the ignition voltage with temperature has the form shown
in Fig. 8.56, which gives the ignition characteristics of a 120 W (5 ft) 'TL'
lamp. We see that as the temperature is decreased from about 5 C, the
ignition voltage increases as the mercury vapour pressure decreases. When
the mercury vapour pressure is practically negligible (below about -30 C),
the rate of increase of the ignition voltage falls off.
LAMP TYPES AND CIRCUITS 167
Ignition
voltage

1~50
!b(\oblt)

-r-..
"'-
1\
~00

350
1\
\ v
300 ~

v
250 \ '/
1\.
'\.... f..-- v
200

150

-50 -so -~o -30 -20 -10 o 10 20 30 ~o so 50


-remp.(c)
Fig. 8.56. Ignition voltage of a 5 ft 120 W 'TL'M lamp as a function of the ambient tem-
perature, on the assumption that the cathodes are always heated to emission temperature.
The use of auto leakage transformers with an open circuit voltage of about
4-5 times the normal burning voltage is an expensive solution involving high
dissipation and a big capacitor to compensate for low power factor. Here
again, semiconductors can be used to give a starter circuit which is switched
off as soon as the lamp has started. The principle of such a circuit is shown
in Fig. 8.57a, while a practical realisation is shown in Fig. 8.57b.
The ignition voltage for the lamp is provided by a small auxiliary trans-
former, while the electrodes are heated with the aid of extra windings. As
soon as the lamp ignites, the current flows through the ballast proper and
the auxiliary capacitor, which is kept small so that the transformer will not
be overloaded. This capacitor current also ensures that the switch of Fig.
8.57a (which is shown as a thyristor circuit in Fig. 8.57b) is closed. The lamp
then burns in series with the stabilisation self-inductance only.

a b

Fig. 8.57. Principle (a) and realisation (b) of circuit for very low temperatures. When the
thyristors are cut off, the transformer provides the ignition voltage. Once the lamp is
burning, the thyristors are triggered; the transformer in parallel with the lamp then ensures
that the cathode voltage is appreciably reduced.
168 FLUORESCENT LAMPS

Example:
One 120 W 'TL'M/RS lamp Vm = 220 V/50 Hz
V1 = 100 V 11 = 15 A
When used at T ~-50 oc:
Unloaded: V 0 = 485 Yeff Vcath = 47 V
Dimensions electronic starter: 44 x 64 x 196 mm
Dimensions ballast: 44 x64 X 128 mm (2 of these ballasts are used in series)
Watt losses: 30 W

8.7 Instant start circuits (cold starting)


As we have explained in Section 8.6, when lamps are ignited without pre-
heating of the electrodes, the open circuit voltage must be made higher than
when the electrodes are pre-heated; see also Fig. 8.39.
This high ignition voltage, which makes the ballast extra large and ex-
pensive, is not needed for the stable burning of the lamp. These circuits often
have two lamps in series, therefore, one of them with a high ohmic impedance
shunted across it. The lamp without the shunt impedance starts at the high
open-circuit voltage, after which steps are taken to ensure that a sufficiently
high voltage is applied across the lamp which still has to be ignited - i.e.
across the auxiliary impedance. These circuits thus make use of the sequence
start principle described in Sections 8.4 and 8.6.
It might be thought that the ballasts for instant start lamps would be
simpler than those for rapid start lamps of the same size and power, since
the former no longer need the windings for heating the cathodes. Unfortu-
nately, this is not the case. If the first lamp ignites at a high open circuit
voltage when the circuit is switched on, this lamp burns in series with the
high impedance shunted across the second lamp.
In Fig. 8.52, for example, the lower lamp gets its current via the auxiliary
capacitance C 1 This very low current (a few milliamperes) is insufficient to
heat the cathodes to emission temperature; an extra voltage drop is therefore
produced to pull the electrons out of the cold cathode. This makes the arc
voltage so high that the voltage remaining across C 1 is insufficient for the
ignition of the second lamp. It is thus necessary to reduce the auxiliary
impedance, so that the current through the lamp which is ignited first is
enough to heat the electrodes to emission temperature. To achieve this, C 1
would have to be drastically increased to at least a few microfarads, depen-
ding on the open circuit voltage and the lamp type. However, such a large
capacitance across a low pressure gas discharge lamp gives impermissibly
high current distortion (see Chapter 7).
This means, therefore, that (expensive) inductances have to be used instead,
giving a circuit of the type shown in Fig. 8.58. The auxiliary impedance in
this circuit is indicated by S 3 ; S 1 and S 2 , together with the capacitor, combine
to give the capacitive sequence start circuit which has already been described;
the two lamps are connected in series across this circuit. For the mechanical
construction of the three windings, see Fig. 8.59.
LAMP TYPES AND CIRCUITS 169

Fig. 8.58. Sequence-start circuit for two Fig. 8.59. Arrangement of the three wind-
lamps without preheated electrodes. The ings of Fig. 8.58. When 4 shunts are inclu-
small capacitance across the lamps which ded, this gives a double auto leakage trans-
can be used in other rapid-start circuits former. The impedance of S 3 is large com-
must be replaced by a bulky auxiliary coil pared with s2.
here.

This circuit works as follows:


In the unloaded state, the voltage across lamp A is Vs 1 + Vs 3 , while that
over lamp B is Vs 3 - Vs 2 ; the former is much greater than the latter (see
the vector diagram of Fig. 8.60). As a result of this, lamp A strikes, and
burns in series with the capacitor and S 3 . The inductive current produced
in this way is so high that the current through lamp A is enough to heat
the electrodes to emission by itself. The series combination of S 1 and S 3
now behaves as an inductively loaded auto leakage transformer; the vector
Vs 3 is caused to rotate by the load on the lamp. (Compare Fig. 8.60b with
Fig. 8.llc of Section 8.4.) The open circuit voltage across lamp B, Vs 3 + Vs 2 ,
increases after the ignition of A until B ignites too. voB in Fig. 8.60b ;?o voB
in Fig. 8.60a. The high ignition voltage, and the need for the auxiliary choke
and magnetic shunts make this ballast quite expensive. Moreover, the ballast

T-----
I
I

:
I

I
I
I
I
I
I
I

:vo
:~ I
I
I I
I
Fig. 8.60. Vector diagram for
I

: l---
I I
\
the circuit of Fig. 8.58. \
\

(a) Directly after switching 1 I

on; since VoA V 08 , lamp ! :voe \,',',,,


A ignites first.
j __ _j_, __
I I

(b)Lamp B can now ignite;


the vector V08 has in-
creased as a result of
rotation of the voltage
vector V83 . b
170 FLUORESCENT LAMPS

hum is generally excessive, as a result of the large size of the ballast and the
magnetic leak fields caused by the four shunts. The use of ballasts of this
type and the corresponding lamps has therefore never been popular in
Europe.
Another solution to this problem of cold starting is provided by the 'TL'S
lamp, whose ignition voltage is of the same order of magnitude as that for
pre-heat starter lamps, despite the fact that the cathodes are not pre-heated.
This means that many of the circuits discussed in Sections 8.4 and 8.6
can be used with practically no modification for 'TL'S lamps. For example,
for the 20 W 'TL'S lamp with mains voltages of 200-250 V:
the inductive circuit is given by Fig. 8.9a (Section 8.4) without the starter;
the capacitive circuit is given by Fig. 8.16a (Section 8.4) without the starter.
At low mains voltages, a step-up transformer or auto leakage transformer
has to be used again.
For 40 W 'TL'S lamps and mains voltages of about 220 V, the semi-
resonance circuit discussed above is very popular; see Fig. 8.45c, Section 8.6.
Of course, the cathode leads are connected in this version. If auto leakage
transformers are used, the open circuit voltage must be at least about 250 V.
If it is not necessary to correct the power factor, the circuit of Fig. 8.61
can be used. A small capacitance (about 08 [LF) together with the normal
switch start choke gives enough voltage step-up to guarantee reliable igni-
tion.

Fig. 8.61. If the open-circuit voltage


required is only slightly higher than
the mains voltage, a small capacitor
(if necessary in series with a small
resistance) may provide the solution
for igniting 'TL'S lamps.

8.8 Lamp operation at higher frequencies


Operation of fluorescent lamps at frequencies considerably above the
normal mains frequencies of 50 or 60 Hz gives a number of striking ad-
vantages:
(1) The ballast can be made much smaller, for the following reasons.
(a) If chokes are used, the same induction is given with fewer windings
this means that the magnetic core can be made smaller too.
(b) The differences in the form of the current and voltage between lamps
burning at different frequencies are illustrated by the oscillograms
of Fig. 7.2, from which it will be seen that, at least for lamps of 1
or 11- in diameter, the form of both current and voltage is nearly
sinusoidal at 400 Hz and higher. This means that capacitors can be
used for stabilisation of the current, without the need for extra self-
inductances in series to limit distortion.
(2) The watt losses will be lower, as a consequence of the points mentioned
in (1) above.
LAMP TYPES AND CIRCUITS 171
% T20

T~~
......
v
7fJW

112
v
/
T08 I v
I
/.. v
T04
~
TOO
5 2 5 2 5 TO'
-t(Hz)
Fig. 8.62. Relative luminous flux (/),. 1 of a 20 W and a 40 W fluorescent lamp, as a func-
tion of the frequency f of the supply current at constant discharge power. r:prel is taken
to be 100% at f =50 Hz.

(3) The lamp efficiency will be higher. Fig. 8.62 shows the relative luminous
flux for 20 W and 40 W lamps as a function of frequency. This increase
in efficiency is due to the reduced anode fall *.
(4) The luminous ripple is very slight, because the afterglow now lasts
longer compared with one cycle of the ripple.
(5) The distortion factor which takes the overall effect of the distortions in
the current and voltage into account (see also Section 7.1), which is
about 09 with normal frequencies (so that e.g. 40 W = 103 V x 043 Ax
09), now becomes nearly 1, so that the lamp current and hence also the
current through the ballast is 10% lower - if the arc voltage is constant.
This is so at frequencies above 15 kHz; but at 400-1 000 Hz the lamp
voltage is appreciably lower than at 50-60 Hz, as may be seen from
Fig. 8.63.
In view of the above-mentioned advantages of operating fluorescent lamps
at higher frequencies, it might be thought that this would be done quite
often. This is not the case, however, because the economics of the system
depend very much on the circumstances.
If the power for the lamps is taken from the normal mains, an extra con-
verter is always needed to give the high frequency voltage (from the rectified
mains voltage). The extra costs and dissipation, not forgetting the special
cables required, make this system too expensive for use for normal domestic
and industrial lighting at present.
Where special demands (e.g. minimum luminous ripple) are made on the
light, or if a d.c. voltage source is available anyway (e.g. in trains and buses),
operation of the lamps at high frequencies with the aid of a d.c.-a.c. inverter
is the optimum solution. This topic is discussed in detail in Chapter 9.
For aircraft lighting, where saving of weight is of prime importance, this
system has obvious advantages - especially as a 400 Hz supply system is
required for other equipment (for the same reasons) anyway.
* SeeM. KoEDAM and W. VERWEY, 'The influence of the supply frequency on the luminous
efficiency of fluorescent lamps', published in: Proceedings of the 7th International Con-
ference on Ionisation phenomenon in gases- Beograd 1965, North-Holland Publishing
Cy., Amsterdam.
172 FLUORESCENT LAMPS

i
%106
101.
102
v ~ I/'
l
I

100
98
l' I ! r--. l
96
'\
'\
91. 1\.
'\. ~
92
90
\ I '
v
v
"\jl I

88 v, v
86 "i..J I I /
81.
10 2 5 5 10 3 2 5 10'
-t(Hz)
Fig. 8.63. Lamp voltage and current for a 40 W lamp as functions of the frequency at
constant lamp power.

Practically all audio frequency fluorescent lighting uses starterless circuits


at present; the circuits discussed in Section 8.6 can be used for this purpose.
As we have seen, there is no need to correct for current distortion; this can
be very useful in certain cases - see e.g. Fig. 8.64, which shows an inductive
and a capacitive circuit in parallel. The lamps ignite immediately the current
is applied, since the L-C circuits are near resonance at that moment, so that
a high voltage is available. Once ignition has taken place, the voltage across
the capacitor or choke in parallel with the lamp adjusts itself to the arc
voltage. In view of the fact that the frequency is relatively high, the capaci-
tors and coils are quite small.

8.9 D.C. circuits and lamps


Direct curreht as a source of power for fluorescent lamps presents some
of the same advantages as high frequency supplies, viz. no luminous ripple
and high lamp efficiency. The great drawback, however, is that the overall
efficiency of the whole lighting installation is much lower than on alternating

Fig. 8.64. Dual-lamp circuit for high


frequencies.
LAMP TYPES AND CIRCUITS 173
current, since the discharge has to be stabilised by means of a resistor which
dissipates about as much power as the lamp itself and in most cases this
power is completely lost. If a tungsten lamp be used as a ballast, the extra
light can of course be put to good use and the efficiency of the whole system
then shows an improvement, but the overall efficiency remains lower than
in the case of operation on a.c. with chokes.
Naturally, a tungsten lamp can be made to function as ballast without any
use being made of its light; a size of lamp can then be selected which will
ensure, on the one hand, an almost unlimited life and, on the other hand,
gives the maximum benefit derived from the good stabilising properties of
the lamp. A tungsten lamp is a good stabiliser, because its character is
strongly positive.
The voltage current characteristic of a gas discharge lamp, however, is
negative, which means that if the current is decreased the arc voltage rises.
Since at a reduced mains voltage the resistance of the filament lamp is
reduced and more voltage is made available for the discharge lamp than
would be the case with a constant resistance, the current therefore diminishes
less than would be the case with a constant series resistor. Fig. 8.65 illus-
trates the difference in sensitivity of the discharge lamp to mains voltage
variations for the two forms of ballast, viz. ohmic resistor and filament lamp.
In the diagram, L and R are the current/voltage characteristics of a gas dis-
charge lamp and constant resistance respectively, when the latter is plotted
downwards from the dotted lines.

---------------~
------------Vm
Fig. 8.65. Stabilisation of a gas dis-
charge lamp by means of either ohmic
resistor or tungsten filament lamp. R
and R' are the current/voltage charac-
teristics of an ohmic resistor at the
mains voltages Vm and Vm'; T and T'
are the corresponding characteristics of
a tungsten lamp. Curve L relates to a
gas discharge lamp, and points S, S' and
S" represent stable working conditions
in which the sum of the voltages on the L
discharge lamp and the resistor (or fila-
ment lamp) is equal to the mains volt-
age.

This method of representation has already been mentioned in Section 1. 7


and the point S indicates the current and voltage at which the discharge is
stable at mains voltage Vm. If the discharge lamp is stabilised by a tungsten
lamp, the latter may be so rated that its characteristic T also passes through S.
It is now assumed that the mains voltage increases from Vm to Vm'; in the
case of the resistor the point of equilibrium then moves to S' (point of inter-
section of L with R'), whereas, with a tungsten lamp as ballast, the point S"
is reached (intersecting point of L and T'). Clearly, then, the current and
voltage variations with filament lamp ballast are much smaller than with
174 FLUORESCENT LAMPS

:~
Fig. 8.66. Circuit for a fluorescent lamp on
220 V d.c. The small inductance L serves
to provide a sufficiently high voltage surge
to ensure ignition when the starter switch
R L S opens.

a constant resistor, for which reason the changes in the consumption of power
and in the luminous flux are also much smaller.
The most popular circuit is shown in Fig. 8.66: a small inductance L is
connected in series with the resistor R to supply the surge necessary for
ignition when the starter S opens. As regards the starter itself, this may be
of the manual type, a thermal switch or a special glow switch for d.c. opera-
tion. This glow switch has to be symmetrical as regards its inner construction,
because it must operate correctly whatever the polarity of the connections.
If the available mains supply is on the low side, a glow switch is unsuitable
and a choice has to be made between the two other types.
All d.c. lamps must be provided with means for reversing the polarity,
to prevent the mercury vapour from being displaced by the electrical field
to the cathode end of the tube (electrophoresis), since this causes the lamp
to give less light at the other end; see Section 2.5.6. Periodic polarity reversal,
that is, every 5 h or less, will prevent this from happening, but it is usually
sufficient to effect reversal automatically each time the lamp is switched on,
with the aid of a special switch.
It is essential that lamps, operating on d.c., are kept at a suitable tem-
perature (approx. 50 oq, because otherwise the vapour pressure at the
cathode end is so low that the diffusion of the mercury atoms, which counter-
acts the electrophoresis, is too small and dark ends will develop fairly soon.
As a rule an enclosed lighting fitting, which at the same time protects the
lamp from draughts, provides a suitable solution. At sufficiently high tem-
perature 11- in lamps up to 2ft do not necessarily require polarity reversal.
The increasing use of fluorescent lamps for lighting ships, trains, tramway
cars, etc. where only d.c. power is available, has led to the development of
some interesting circuits and equipment.
Because of the special character of the above applications, where great
reliability and simplicity of installation are required, solutions which ob-
viated the use of a starter switch were favoured. As a typical example we
mention the system which was adopted for the lighting of trains where a
d.c. supply of 72 V is available. The circuit chosen is depicted in Fig. 8.67.

+
Fig. 8.67. Circuit for low d.c. voltages
(72-100 V).
LAMP TYPES AND CIRCUITS 175

Fig. 8.68. Circuit for 'TL'C lamps


with a relay which interrupts the
pre-heating current as soon as the
+-:1@
_
C---~
_

lamp has struck.

A stabiliser tube containing two resistors, R 1 and R 2 , is connected in series


with a 'TL'C 15 W lamp (18 x 11 in); these lamps, and other special d.c.
lamps, have a special electrode design to avoid end blackening. The circuit
operates as follows: Switches no. I and 2 are closed, cathode A is then pre-
heated and the lamp strikes on the supply voltage which is available between
A and B. After striking, switch no. I is opened in order to interrupt the
heating current. The lamp current is limited by resistor R 2 , which has sta-
bilising characteristics similar to those of a filament lamp. R 1 limits the pre-
heat current through the cathode.
If the supply voltage is in the vicinity of 100 V, the 'TL'C 20 W lamp is
the most suitable type. This lamp is provided with two inside strips, exten-
ding over about half the lamp length. Each strip is connected with one end
to an electrode, their other ends are about 1 in apart near the middle of the
lamp. This construction has resulted in a very low ignition voltage, which
makes these lamps suitable for use on supplies with very large fluctuations
in voltage, such as are encountered in train lighting.
Instead of the use of push buttons or other manual switches, as shown in
Fig. 8.67, a relay can be used to cut out the pre-heat current as soon as the
lamp has struck (see Fig. 8.68). The lamp current flows through the relay
coil and actuates the switch, which cuts off the current through the lamp
cathode.
Since the stability of a discharge is better on d.c. than on a.c., because no
periodic re-ignition of the arc is necessary, it has been possible to choose a
lamp with an arc voltage which is relatively high with regard to the supply
voltage, thus minimising the losses in the stabilising resistor. The overall
efficiency is therefore quite acceptable.
Where d.c. supplies of several hundred volts are available, as in the case
of tramways, starterless operation is possible with lamps of the 'TL'R type.
These are derived from 'TL'S lamps by applying a second conducting strip
along the inside of the tube, this one connected to the other electrode. This
makes the lamp independent of the polarity of the supply, which is necessary
because of the required reversal of polarity.
On 450-650 V supplies 2 'TL'R 40 W lamps can be operated in series,
with an ample margin for voltage drops, which frequently occur in tramway
supplies. For stabilising the discharge, incandescent lamps or a fixed resistor
of a suitable value are commonly used. In case incandescent lamps are
employed, these may serve usefully for lighting route indicators, line numbers
or the like.
Instead of two 'TL'R 40 W lamps, three 'TL'R 20 W lamps may be used
in series on 450-650 V supplies. To facilitate striking, one of these lamps
should be bridged by a (wirewound) resistor of about 10 Q.
176 FLUORESCENT LAMPS

Fig. 8.69. Circuit for 2 'TL'R 20 W lamps in series with relay for sequence-starting of the
lamps.

For ship lighting on 220 V d.c. supplies the circuit of Fig. 8.69 for 2 'TL'R
20 W lamps in series, with stabilising tube and relay, has proved to be a
practical proposition.
If polarity reversal is not required, which may be the case under favourable
temperature conditions and with relatively short lamps, 'TL'S 20 W lamps
can be used. The lamp must then be connected in series in such a way that
for all three lamps the electrode not connected to the strip is connected to
the negative side of the supply (see Fig. 8. 70).
R'

+'"'
r _____
rr------------
\ ,' r--------------
--IJI----1\1--
' r-------------
-Jf---_.__-+\~
'
.J

Fig. 8. 70. Series circuit for 3 'TL'S lamps.

In this circuit R represents the stabilising resistor. R' is a resistor across


one of the lamps, the other two will start first, followed by the third one.
This arrangement is necessary because the combined striking voltage of the
three lamps in series is greater than the available supply voltage.
On 220 V d.c. mains a 'TL'R 40 W lamp can be operated in series with a
110 V 40 W incandescent lamp. A resistor of a constant value is less suitable
in this case. Striking characteristics and regulation are much better with an
incandescent lamp or a stabilising tube.
The advent of transistor and thyristor circuits has reduced the relative
importance of d.c. operation of fluorescent lamps, particularly for supply
voltages below 120 V (see Chapter 9).
To conclude this section we give in Table 8.6 some data regarding special
d.c.lamps.
LAMP TYPES AND CIRCUITS 177
Table 8.6 Data of special d.c. lamps
approx.
lamp type dimen- cap arc current lumens (Cool
sions voltage White)

'TL' C 15 W 18xltin bi-pin 45 033 A 800 lm


'TL'C 20 W 24 x It in bi-pin 56 036 A I 100 lm
'TL'R 20 W 24x It in recessed single contact 59 034 A 900 lm
'TL'R 40 W 48x It in recessed single contact ll5 035 A 2 300 lm

8.10 Lamp types for special purposes


The ever increasing range of fluorescent lamps has d~veloped more or less
along the same lines as for incandescent lamps. After the normal lamps for
general lighting purposes various kinds of special lamps appeared, e.g.
reflector lamps, coloured lamps and lamps radiating in certain regions of
the ultra-violet spectrum.
The fluorescent reflector type lamp * (Philips 'TL'F) is a useful tool for
the illuminating engineer. Its advantages are that the intensity distribution
of the light is asymmetrical and that the light output is only very slightly
influenced by dust falling on the lamps. The principle of the lamp is shown
in Fig. 8.71.
About i of the circumference of the lamp is coated on the inside with a
white reflecting layer. The normal fluorescent coating is applied over this
reflecting layer and the clear window. Extensive research on the composition
of the reflecting layer and the width of the window has resulted in a lamp
giving about 90% of the lumen output of a normal lamp, but with an inten-
sity in the direction of the window of about 18 times that of the normal lamp.

Fig. 8.71. Schematic cross-section of a 'TL' F lamp.


The portion S, subtending an angle 2nj; is coated
only with a norma l fluorescent powder layer. Over
the rest of the circumference, i.e. the portion R, a
layer of powder of good reflection properties is applied
between the fluorescent powder and the glass wall.
The portion R has thus a higher reflection factor and
also a greater absorption factor than the portion S.
H represents the luminous intensity vertically down-
wards and It the luminous intensity upwards.

* ]. J. BALDER and M. H. A. v .o. WEYER, Philips Techn . R e v. 17, 198. (1956)


178 FLUORESCENT LAMPS

Fig. 8. 72. Light distribution cur-


ve of the 40 W 'TL'F lamp. For
comparison purposes the corres-
ponding curve for a normal lamp
is indicated with a dotted line.

The intensity distribution is given in Fig. 8.72. The brightness of the window
is of course fairly high, the lamps should therefore be used in fittings provi-
ding sufficient shielding. These fittings may be of a simple design which
adds to the economical advantages of the fluorescent reflector lamps.
A comparison of the lumen maintenance of these lamps and of normal
lamps in fittings shows clearly the superiority of the reflector lamp. The
efficiency of normal lamps and fittings is considerably reduced by dust sett-
ling on the lamps and on the inner surface of the fittings. Since the reflector
lamps give only a small fraction of their total lumens in the upper hemisphere,
the influence of dirt is here very small. 'TL'F lamps are manufactured in
20-40-65 and 80 W sizes, in various colours.
Coloured lamps are made by using special phosphors, if necessary in com-
bination with a lacquer on the outer surface of the tube or a pigment added
to the phosphor on the inside. Red, green and blue are the main colours.
Lamps for the ultra-violet region can be split up in the following cate-
gories:
(l) Germicidal lamps. These lamps are not provided with a fluorescent
coating, but since they have the same electrical characteristics as normal
fluorescent lamps, they are treated here. The bulb is made of special
glass, which transmits a large part of the 254 nm radiation, generated
in the low pressure mercury discharge. These lamps, the radiation of
which is capable of killing bacteria and germs, have found widespread
application in the food and pharmaceutical industry, and also in hos-
pitals.
(2) Black light lamps, radiating in the region around 360 nm. This radiation
may be used to obtain fluorescent effects e.g. for advertising; it can also
LAMP TYPES AND CIRCUITS 179

0/o

liN
TOO

I I
I I

1 I ~
80

I \

I
60
I I
fS;..
ll\
~

I
I

II
40
I
I
I
I

IJJ n
20

n
\ ~
' .... 1\
0
250 300 350 400 "' 450i'-- 500 550 600mfl
Fig. 8.73. Relative spectral energy distribution of fluorescent photocopying lamp (E;)
and spectral sensitivity of average diazo-printing paper (S;).

be used for printing purposes. When fluorescent effects are to be obtained,


it may be that the visible light of the lamp is troublesome. In those cases
a lamp with a dark blue bulb should be used. These 'black light blue'
lamps radiate almost exclusively in the ultra-violet region. They are ap-
plied for chemical analysis or detection of small traces of organic subs-
tances.
(3) Photocopying lamps, using a phosphor with a maximum in the radiated
energy near 400 nm. This coincides reasonably well with the maximum
of the sensitivity curve of average diazo-printing paper. See Fig. 8.73.
In recent years remarkable progress has been made in the development
of phosphors for this application. The output of the lamps has been
increased and at the same time the influence of temperature on the
efficiency of the phosphors has been considerably reduced. For use in
xerographic reproduction equipment a special lamp in the dimensions
of the lin 15 W lamp, but with a higher loading, has been developed.
The phosphor emits in the green part of the visible spectrum and an extra
high intensity is obtained by applying an internal reflector ('TL'F prin-
ciple) with a narrow aperture.
180 FLUORESCENT LAMPS

The various types of lamps with ultra-violet radiation are manufactured


only in a limited number of sizes. A rough idea of the available types is
given in the summary below.
Blacklight lamps 15-20-30-40-65 W
Blacklight blue lamps 20-40 W
Printing lamps 15-25-40-65-80-120 W
Germicidal lamps 15 and 30 W.
Chapter 9

Inverters and converters


Th. Hehenkamp

9.1 Introduction
The advantages of lamp operation on higher frequencies than 50 or 60 Hz
have already been pointed out in Section 8.8. These advantages are: smaller
ballasts with lower losses, a higher lamp efficiency and a negligible luminous
ripple. Several attempts have been made to make use of these advantages in
practice. This is not too difficult in the case of installations for which some
form of inversion is needed anyway, as for example where only a low voltage
d.c. supply is available. A d.c.-a.c. inverter can then be employed to provide
the optimum value of output frequency. Values between 100Hz and 25kHz
are in use; the higher frequencies exploit more fully the possibilities referred
to above, but are more difficult to achieve with a high inversion efficiency.
For installations operating from the normal 50 or 60 Hz power supplies it
is more difficult to obtain favourable results. The necessity of an a.c.-d.c.
converter, introducing higher investments and appreciable losses, in most
instances more than counterbalances the possible gain.

To obtain satisfactory results the inverter or converter must meet the


following requirements:
1. Reliable operation, without maintenance
2. A high output frequency
3. A high efficiency
4. Low cost
182 FLUORESCENT LAMPS

These conditions are best met by circuits using semiconductors. Such cir-
cuits require no servicing and can have a high operating frequency and a
high efficiency. Other solutions such as rotary converters, mercury-jet invert-
ers and magnetic converters are less satisfactory and will have been ousted
by the semiconductor circuits within a few years' time.
Transistors are used for low wattages, thyristors for higher wattages, both
in inverter circuits. If an a.c. supply is available the inverter is extended into
a converter by including a rectifier with silicon diodes.

9.2 Principle of the transistor inverter


Fig. 9.1 shows a simple circuit for an inverter, in which the load is repre-
sented by the resistance R 0 This is basically the well-known vibrator circuit
in which the two switching contacts have each been replaced by a transistor
and which has control windings for the transistor bases. The transistors
operate as switches * and alternate in passing current from emitter to col-
lector and the transformer voltage is practically square wave in form. The
fact that a relatively large power can be switched with the help of transistors

Tr

Q.

Fig. 9.1. (a) Simple circuit diagram for a DC-AC inverter with transistors, loaded with a
resistance R 0
(b) The transformer Tr and the resistance R 0 are replaced by the parallel combination
of a self-inductance L and resistance R.
(c) The two components (iR and iL) of the collector current ic.

* R. L. BRIGHT, 'Junction transistors used as switches', AlEE Trans. 74, pt I, 111-121


(1955).
INVERTERS AND CONVERTERS 183
is due to the circumstance that only a small voltage loss occurs over the
conducting transistor, whilst only a very small leakage current flows in the
blocking transistor, although a considerable voltage can be present across
it. The combination of a high current through and a high voltage over the
transistor occurs only during switching and lasts for only a small part of the
cycle, at least at low operating frequencies. The circuit of Fig. 9.1 operates
as follows: If transistor T1 is conducting and T 2 non-conducting, the battery
supplies current via the emitter and collector of T1 , this current passing
through w1 As long as this is the case, practically the whole of the battery
voltage Vb is across w1 , whilst constant voltages are also present across w3 ,
w4 and w5 Now w3 and w4 are so connected that the base of T1 is made
more strongly negative than already achieved by the potential divider R 1-R 2 ,
and that the base of T 2 is positive. This is in conformance with the conducting
condition of T1 and the blocking condition of T 2
The coils w1 and w 5 together with resistance R 0 can be replaced by a
resistance Rand a self-inductance L connected in parallel (Fig. 9.1b). Over
this parallel circuit is the voltage Vb (if we neglect the loss in T1 ). The col-
lector current ic is then divided into two components: the constant current
iR = Vb/R and the current, increasing in direct proportion to the time,
iL = Vb t/L. L is assumed here to be constant. The collector current
increases thus with the time. This cannot proceed indefinitely, however,
since an upper limit Ic is set to ic, this being determined by the constant
voltage over w3 As soon as ic has reached this value a further increase
becomes impossible. This also applies to the component iu so that diL/dt
becomes zero and the transformer voltages fall away. The voltage across w3
thus disappears, so that the base current of T1 decreases and ic must also
decrease. As a result, voltages are now induced in the transformer windings
with the opposite sign to the original voltages. In consequence no further
current passes through T 1 and T 2 becomes conducting: T 2 and w4 now come
into the condition previously applying to T 1 and w3 A periodically alterna-
ting condition is thus set up. One half cycle is (neglecting the extremely short
switching-over period) equal to the time which ic needs to increase from the
value Vb/R to the value Ic. It is thus easy to see that this time will be pro-
portionally shorter (and the frequency thus higher) as L, Rand Ic are made
smaller and Vb greater.
The load resistance in Fig. 9.1 cannot be replaced directly by a fluorescent
lamp. This is a consequence of the negative current-voltage characteristic.
As soon as the lamp ignites, the momentary value of the collector current
through the then conductive transistor increases rapidly and reaches the
value Ic very quickly. The operating frequency of the inverter will conse-
quently be much higher than if resistance loading is used and will be so high
that switching losses of the transistor will start to play a dominant part.
To counteract this a delaying element must be introduced in series with
the lamp, a self-inductance being quite suitable for this purpose. The fre-
quency may then be set for a few kHz, proper operation being realised at
an efficiency of 50-60%. To attain a higher operating frequency and/or a
higher efficiency, it is necessary to change the circuit so that the switching
losses become less substantial.
184 FLUORESCENT LAMPS

9.3 Inverter with reduced switching losses


The collector current through a transistor increases or decreases more
slowly due to physical causes than is indicated by the basic current, par-
ticularly at high currents this inertia may be considerable (Fig. 9.2a). This
has its consequences every time the transistor becomes conductive or switches
off. When the transistor becomes conductive the collector current will in-
crease less rapidly than desirable so that less energy is supplied to the load.
An improvement is realised by giving the base current a high starting value
(Fig. 9.2b).

... -- ---

Fig. 9.2. ic = f(ib) of a transistor,


(a) with constant base-current pulse;
(b) with basecurrent pulse of high initial
value, the direction of the base current
being reversed at the moment when the
transistor is switched off, so that ic rises
and falls faster than in Fig. 9.2a.

When switching off the collector current decreases relatively slowly. In


the circuit of Fig. 9.1, a high blocking voltage will be applied across the
transistor at the moment when the collector current starts to decrease due
to the action of the inductance, and the time integral of the product of this
voltage and the decreasing collector current can achieve considerable values.
This can be improved by changing the direction of the base current at the
moment of switch-off, the residual charge in the base region then being
removed more rapidly (Fig. 9.2b).
INVERTERS AND CONVERTERS 185
It is even more effective to have the blocking voltage across the transistor
rise slowly from zero by having a capacitor in parallel to the transformer.
This means that the second transistor can only he made conductive after a
short delay, the transformer voltage no longer being square-wave but rather
trapezoid (Fig. 9.3).
In the circuit shown in Fig. 9.4 all these steps have been taken to reduce
the effects of inertia of the transistor *.

Fig. 9.3. Trapezoid transfor- l.'tr


mer voltage, produced when
a capacitor is switched in
parallel with the transformer
by means of a transistor
inverter. The transistors con-
duct in turn, in the periods
t 1 -t 2 and t 3 -t4

Fig. 9.4. Simple circuit diagram of a transistor inverter with base-current pulse like that
of Fig. 9.2b, with a network (L, and C 3 ) for production of a sinusoidal transformer
voltage.

Moreover, the small series inductance L, has been provided which gives
the transformer voltage a smoother shape so that it can become more
sinusoidal than shown in Fig. 9.3.
Capacitors C 1 and C2 provide pulse-shaped basic currents and also ensure
that the base of the transistor concerned has a high positive blocking voltage
relative to the emitter at the moment of switch-off. Fig. 9.5 shows the results
of these measures to increase the switching speed of the transistor. It has
been found possible to achieve efficiencies of 70~80% at operating frequen-
cies of 15 kHz and higher.
* I. F. DAVIES and D. DuNTHORNE, 'The application of power transistors to the operation
of gas-discharge lamps from d.c. supplies'. Proc. I.E.E., 107, Part A, No. 33, 273 (1960).
J. J. WILTING, 'Die Entwicklung auf dem Gebiete der Transistorumformer fiir Fluorenz-
lampen', Bulletin S.E. V., 53, No. 22, 1082-1091 (1962).
186 FLUORESCENT LAMPS

Fig. 9.5. Top : The collector currents of the two transistors of the inver-
ter of Fig. 9.4. The form of this current corresponds basically to that
of Fig. 9.2b, but the effect of Ls and C 3 gives rise to an extra oscillation
which causes the rounding-off of the top of the pulse, and the form
shown for the training edge.
Bottom: The transformer voltage (roughly sinusoidal) of the inverter
of Fig. 9.4.

9.4 Cooling of the transistor


The transistor proper is formed by a plate of germanium or silicon of only
a few square mm in area and a few tenths of one mm thickness. Although
the transistor is a satisfactory switch, which has relatively few losses owing
to the steps just described, the specific load is tremendous due to its small

Fig. 9.6. Pressure-diecast frame with the two transistors of a transistor inverter mounted
on it. The six lugs shown ensure good thermal contact between the transistors and a heat
sink.
INVERTERS AND CONVERTERS 187
dimensions. That is why the crystal has beep secured in perfect thermal con-
tact to a copper base which in turn must have a very good thermal contact
with a heat-sink. A suitable method to do so is to have a transistor inverter
with a thick-walled aluminium frame to which the transistors are secured
in good thermal contact and which also carries the other components. This
frame can have so much cooling area that the transistor temperature is kept
within permissible limits if the ambient temperature is moderate. If the
inverter is built into a small space that hampers cooling and/or used at a
high ambient temperature, the inverter must be secured to a metal surface
that permits sufficient heat dissipation. A construction that has a particularly
low internal temperature drop is to be preferred in this case, e.g. a frame of
pressure-diecast aluminium (Fig. 9.6).

9.5 Applications of transistor inverters


As stated in the introduction to this section, the use of transistorised
inverters is logical in cases where a fluorescent lamp has to be supplied from
a battery. Admittedly, if the battery voltage exceeds approx. 70 V it is pos-
sible to connect short fluorescent lamps to the battery through a resistance
that has current-controlling features, but this has the drawback of a modest
efficiency and the risk of mercury displacement due to electrophoresis and
by consequence a necessity of regularly changing the polarity. At lower
battery voltages this system is no longer applicable and the voltage applied
will have to be increased in some way or other. Battery supply is usual
in aircraft, vehicles and vessels and in emergency lighting of buildings. The
voltage used varies between 6 and 120 V, the lamp power required usually
varies between 6 and 40 W. This range fits very well to the possibilities of the
transistor, germanium transistors being normally used for the lower voltages
and silicon transistors for the higher voltages. In some cases a considerable
power is required to feed a group of lamps from a central inverter. For
this purpose the transistorised inverters are less suitable, because transis-
tors for high currents which are also fast are hard to make. In such cases
the use of thyristors is preferred.
Figs 9.7, 9.8 and 9.9 show some examples of the use oftransistorinverters.

9.6 The thyristor as a switch


A thyristor is a semiconductor with three junctions and characteristics
much like those of a thyratron. This means that without any suitable signal
applied to the control electrode (the gate), the thyristor remains blocked,
irrespective if the anode is positive or negative relatively to the cathode.
However, if the gate is made sufficiently positive relative to the cathode, the
thyristor becomes conductive, provided the anode is also positive* (Fig. 9.10).
The conductive state then no longer depends on the control signal. This may
be interrupted or changed in polarity without any effect. The conducting
* I.M. MACKINTOSH, 'The electrical characteristics of silicon PNPN triodes', Proc.
IRE, 46, 1229 (1958).
188 FLUORESCENT LAMPS

Fig. 9.7. Lighting of an aircraft cabin (a Caravelle of Air France) by'TL' lamps with power
supply via transistor inverters.
INVERTERS AND CONVERTERS 189

Fig. 9.8. Lighting of a bus with 'TL' lamps. Supply voltage 24 V.

Fig. 9.9. Example of a 'TL' lighting system for caravans, boats and tents.
190 FLUORESCENT LAMPS

Fig. 9.10. vi.- Vac characteristics of a


thyristor.
(1) cut-off characteristic in reverse direction
(2) cut-off characteristic in forward direc-
tion
(3) characteristic in forward direction, after
the thyristor has been made conducting by
a gate pulse
(4) making a thyristor conducting by in-
creasing the forward voltage above the cut-
off value; this effect is often undesirable.

state can only be terminated by reducing the anode current until it equals
or is less than a specific low value called the holding current. In inverter
circuits the current is usually interrupted by sending a counter current
through the thyristor. In that case a blocking voltage will occur across the
thyristor, renewed conductivity only becoming possible after the blocking
voltage has been replaced by a forward voltage and a suitable signal has
been applied to the gate.
This means that the characteristics of a thyristor and a thyratron are very
similar although there are substantial differences, which are all advantages
for the thyristor:
(1) the dimensions are much smaller;
(2) no heating of the cathode is required;
(3) the potential drop in the forward direction is small, usually between 05
and 2 V;
(4) the switching speed is much higher.
All these points render the thyristor very suitable for operation as a switch
in inverter circuits with operating frequencies up to 1 kHz. For higher fre-
quencies the switching speed is often on the low side and this requires special
care. The first problem is a result of the relatively low turn-on speed, the
main current starts in the gate region and lasts for a few [lS before the cur-
rent has distributed uniformly throughout the crystal. With a steeply in-
creasing current the resulting high initial value of the current density in the
thyristor causes a much higher voltage drop than the above 05-2 V, and
consequently there will be a considerable loss of energy. This loss can be
restricted by having the current increase gradually after switching on, an
inductance in series being a suitable expedient (Fig. 9.ll).
Switching off, as stated earlier, can be effected by lowering the anode cur-
rent down to the holding current, but in that case the turn-off time will be
quite long and in the order of 100 [lS. A higher speed can be achieved by
applying a blocking voltage across the thyristor to send a counter-current
through it and so forcing the current to zero. In that case there will be two
successive states to consider.
INVERTERS AND CONVERTERS 191

i
r------
1
I
I
I

Fig. 9.1 I. Switching on of a thyristor. Use of delay elements means


that the current i in 9.11b increases more slowly than in 9.1la, so that
the voltage v in 9.11 breaches its final value more quickly. The switching
losses are thus lower in the case of 9.11b than in that of 9.11a.

At the moment when the current becomes zero, there are still so many
charge carriers near the outer junctions that the current will momentarily run
in the opposite direction, the so-called recovery current (Fig. 9.12). As soon
as the residual charge has been removed, the recovery current will break off
abruptly and a reverse voltage can then appear over the thyristor because it
is now able to block this. However, as long as charge carriers exist in the
vicinity of the centre junction, the thyristor can not block a forward voltage,
the conductive state returning as soon as a forward voltage is applied. So a
short turn-off time must elapse during which the remaining charge carriers
can disappear due to recombination. The turn-off time in fast thyristors is
only a few [LS but normally 10-20 [LS and even more is required.
The fast break-down of the recovery current can lead to a high L di/dt
value if an inductance has been placed in series with the thyristor. So a high
voltage will appear over the thyristor and damage may result. In such a case
an R-C circuit must be provided in parallel to the thyristor to reduce the
peak voltage to an acceptable value. If due attention is paid to all the points
mentioned, it will be possible to raise the operation frequency of a thyristor
inverter to approx. 10 kHz. Still higher values will only be possible with very
special circuits which are outside the scope of this book.

Fig. 9.12. Typical waveforms of thyristor


I voltage v and current i. The part of the
I current below the zero line is called the
I recovery current. After switching off (time
I
I t 1 ), the voltage should be negative until t 2 ,
I so that the thyristor can become cut off in
I the forward direction again. 12 - t 1 = t 0
I~
is called the turn-off time of the circuit;
this should exceed the turn-off time of the
thyristor.
192 FLUORESCENT LAMPS

9.7 Inverter with forced commutation


Section 9.2 discussed a transistor inverter derived from the well-known
vibrator circuit by replacing the vibrator contacts by transistors and applying
a suitable base control. In approximately the same way a thyristor inverter
can be made using thyristors instead of transistors and adapting the control
circuit of the two thyristors. Since the control circuit can only make the
thyristors conductive but can not switch them off, a commutation capacitor C
and an inductance L have been added (Fig. 9.13) *.

r,

L
c +

Fig. 9.13. Basic circuit diagram of a thyristor inverter with forced com-
mutation.

The operation of the two semiconductor inverters will therefore show


distinct differences: in the transistor type switching from one transistor to
the other is achieved because the collector current of the conductive tran-
sistor reaches at a certain moment a limit set by the base current available
and will not increase any further, the decreasing current change causing an
induction voltage which blocks the transistor. At the same time or a moment
later the other transistor is made conductive. The switching signal therefore
comes from the conductive transistor, which blocks itself- an effect that
can be used to make the circuit self-oscillating. In the corresponding thyristor
inverter the conductive thyristor does not give the signal itself because the
current increase is not limited at all so that an external signal must block the
conductive thyristor at the end of each half cycle. The simplest way is to
derive this signal from a separate control generator: the inverter is then not
self-oscillating but has a forced frequency.
Let us assume (Fig. 9.13) that thyristor T 1 is conductive and consequently
commutation capacitor Cis charged to double the supply voltage. Now the
non-conductive thyristor T 2 receives a signal from the control generator
which makes it conductive also. Consequently the capacitor will now be in
parallel to T 1 and will discharge over this so that the current through T 1
is lowered, changes its direction and T 1 switches off. The capacitor is now
* C. F. WAGNER, 'Parallel inverter with resistance load', Electrical Engineering, 1227-1235
(Nov. 1935).
INVERTERS AND CONVERTERS 193
---,Vr Vrr

Fig. 9.14. Thyristor voltage VT and transformer voltage V,, of the cir-
cuit of Fig. 9.13. The negative part of VT (in the period from t 1 to t 2 ) is
needed for turning the thyristor off; t 1 - t 2 is therefore called the turn-
off time.

charged in the opposite direction through T 2 by the battery. Inductance L


limits the amplitude of battery current and thyristor current during com-
mutation and prolongs their rise time. Owing to this delay T 1 keeps a block-
ing voltage for a sufficient time (Fig. 9.14). At a low operating frequency
the transformer voltage is practically rectangular in shape, at a higher fre-
quency it will be slightly trapezoid to give sufficient turn off time.
Such an inverter will be very suitable for frequencies to approx. I kHz
and for fixed resistance loads. A variable resistance load will entail the
problem of the commutation capacitor having to be large enough to com-
mutate the thyristor current at maximum load. At a smaller load, or worse,
at zero load the capacitor will be too large so that high switching voltages
occur which may jeopardise the thyristors and change the output voltage
substantially. These aspects will be even worse if the load is inductive *,
which is hard to avoid in loads consisting of fluorescent lamps because each
lamp must have a current-stabilising impedance in series. Various circuits
are known which cope with these problems in some way or other, resulting
in satisfactory inverters for low operating frequencies**. However,for fre-
quencies exceeding I kHz a different type of inverter has been found more
suitable: one commutating itself by means of an oscillatory circuit.

9.8 Self-commutating inverters


If the inverter is designed as a properly chosen resonant circuit which is
under-critically damped by the load, a damped oscillatory current will tend
to flow from the supply to the resonant circuit and load when a thyristor
is made conductive (Fig. 9.15). The first time this current passes through zero
the thyristor will automatically turn off, whence the name of self-commuta-
ting inverter. A second thyristor can deal with the next half cycle, being made
conductive just after the turn-off of the first thyristor. This can be done by
means of a separate control generator, but also by deriving a control signal
from the switch-off process of the first thyristor and applying this to the
second thyristor with a suitable delay time. We then have a self-oscillating
* C. F. WAGNER, 'Parallel inverter with inductive load', Electrical Engineering, 970-980
(Sept. 1936).
** B. D. BEDFORD and R. G. Hon, Principles of inverter circuits, John Wiley & Sons,
New York (1964).
194 FLUORESCENT LAMPS

t ~R
.i T
.......
L
'OWOO' I
,,\ .
II \\ _ _,- I - -11

:.
1 .l-- I
I .,.. "" \ I
y...,. \ I
..-'I '-'
Q ..-"' I
I
I
I JL
1\ I
t, .._1
Fig. 9.15. (a) Switching on of an under-critically damped circuit at the voltage V via a
thyristor T.
(b) Wave form of the resultant current ir, consisting of an exponentially increasing por-
tion i 1 and a damped sinusoidal component. The current ir flows from t 0 to t 1 ; at t~>
the thyristor is cut off by the reversal of the current.

circuit which shows a likeness to the transistor inverter of Section 9.3.


Various forms of resonant circuit can be used, such as a series circuit of
inductance and capacitance in series with the load (series inverter) or with
the load in parallel to the capacitor (parallel inverter)*, or with more
complicated circuits which give a more sinusoidal output voltage**.
In several of these circuits diodes are connected in parallel with the thy-
ristors to feed back reactive power from the circuit and the load to the
mains so that the output voltage of the inverter becomes less dependent on
the load.
A variation of the parallel inverter has been found very attractive for use
with fluorescent lamps (Fig. 9.16) ***.The operation is as explained below.
If the thyristor is made conducting at to (Fig. 9.17) a current consisting
of two components begins to flow. One component increases in direct pro-
portion to the time, the other is sinusoidal.
Provided that the values of L1o L 2 and C are chosen appropriately the
resulting current will have a zero transition at time t 1 and the thyristor will
then become non-conducting. Between t0 and t 1 the thyristor conducts and
energy is supplied to C and L 2 From t 1 onwards the circuits LrC oscillates
at its resonant frequency in a damped sinusoidal manner. For a short time
after t 1 the voltage across the thyristor is in the reversed direction, t 2 being
the time at which the voltage changes again to the forward direction. The
interval t 2 -t 1 is the time interval available to turn off the thyristor.
From t 2 to t 3 , the end of the first half period, the thyristor voltage is in
the forward direction. In the second half period from t 3 to t 4 there is only
a sinusoidal voltage. Thus the output voltage of a one thyristor inverter is
asymmetric. With a push-pull arrangement the second thyristor is made
conducting at about t 3 , a symmetric output voltage thus being obtained.
Since during the state of equilibrium attained after a few cycles the mains
* J. J. WILTING, 'D.C./A.C. converter using silicon controlled rectifiers for fluorescent
lighting', Philips Techn Rev .. , 23, no. 8/9, 272 (1961-62).
**R. R. Orr and L. A. SCHLABACH, 'A unique silicon controlled rectifier high power
inverter with sine-wave output voltage', Digest of technical papers 1962, International
Solid-State Circuits Conference, pp. 100--101.
u* TH. HEHENKAMP. 'Thyristor inverters for fluorescent lighting' Trans. Ilium. Eng. Soc
32, no. 4, 197-206 (1967).
INVERTERS AND CONVERTERS 195
T L,

Fig. 9.16. Basic circuit diagram of a self-


commutating one-thyristor inverter.

Fig. 9.17. Voltage and current waveforms


of the circuit of Fig. 9.16.

only has to supply energy to make up for the losses of the unloaded inverter,
each thyristor is only made conductive when the forward voltage has fallen
to a very low value. The current pulse will then be small and of short duration
(Fig. 9.18).
To ensure stability in operation it is necessary that the pulses that render
the thyristors alternately conductive should be supplied at exactly the right
instant, a slight error in the repetition frequency giving a large difference
in output voltages. This precision can be achieved readily by deriving the
pulses from the inverter action itself and thus making this self-oscillating.

a b
Fig. 9.18. Thyristor current (top curves) and transformer voltage (bottom curves) of a
two-thyristor inverter on the principle of Fig. 9.16. The currents of the two thyristors
are shown:
a) in the unloaded case;
b) loaded with a number of 'TL' lamps, each with a series capacitor as ballast.
The two graphs are on the same scale; the loading causes the transformer voltage to
depart from a sinusoidal wave form.
196 FLUORESCENT LAMPS

....

IL
Fig. 9.19. Lighting installation in a climatic test room for botanical experiments.

The inverter described is unsuitable for pure resistance loading. Such loading
will have a damping effect on the oscillatory circuit so that the voltage across
the thyristor is lowered at the moment when it becomes conductive and
consequently so is the thyristor current which ought to rise in order to supply
energy to the load. Loading of a capacitive nature is necessary for proper
operation so that as the load increases the increasing damping is compensated
by an adaption of the capacity on the oscillatory circuit. This is ideal for
fluorescent lamps because the load may consist of lamps that each have a
capacitance in series as a ballast.
9.9 Applications of thyristor inverters and converters
As stated earlier, thyristor inverters are most suitable where transistor
inverters cannot supply the power required. This is the case in central invert-
ers for vehicles and vessels that carry their own supply. The most common
voltages are between 36 V and 120 V, values which are quite suitable, par-
ticularly at the higher voltages efficient use can be made of thyristors. The
ballast is usually fitted close to the lamp whereas inverters are usually
installed at a central point, where effective cooling is possible.
Other uses may be found by adding a rectifier to the inverter to rectify the
normal a.c. supply installed in a building. We then get a converter circuit
which converts the 50 or 60 Hz voltage into a voltage of a much higher fre-
INVERTERS AND CONVERTERS 197

Fig. 9.20. Power-supply unit (con-


sisting of I 2 thyristor inverters) for
the lighting in two climatic test
rooms as shown in Fig. 9.19.
quency intended to exploit the advantages of the higher frequency. The com-
plete absence of stroboscopic effect is an important advantage in lighting
machines for plastic-fibre manufacture and in spinning-mills. The high effi-
ciency of the lamps and the small size of the ballasts are advantages in the
lighting of climatic test rooms containing plants under controlled conditions.
A very high illumination is wanted there with a minimum of room for lamps
and ballasts (Fig. 9.19). The general lighting of factories and offices can also
be designed in this way, the high efficiency of the lamps and the low losses
of the ballasts then being the major advantage. The small dimensions and
the low weight of the ballast permit a lighter construction of fittings and
ceilings, which helps to compensate the high investments for the converter
installation.
Inverters and rectifiers are best installed in a separate control room, which
practice has the added advantage that the heat developed by this apparatus
is being kept outside the room lighted. In this way it will be possible to
economise handsomely on air-conditioning expenses in air-conditioned
buildings, but in spite of this advantage a fluorescent lighting using thyristor
inverters is usually not yet justified economically.
Chapter 10

Dimming of fluorescent lamps


J. C. Moerkens

10.1 Introduction
Now that controlled semiconductor devices are available, modern lighting
control equipment can be made simpler and smaller than was possible when
the only control devices were thyratrons and variable resistors or transfor-
mers. Semiconductors are being used more and more where continuously
variable lighting is required - not only indoors (e.g. in theatres, lecture
rooms, etc.) but also out of doors (e.g. in road tunnels). In factories too,
where a constant lighting level is to be preferred, the fluctuating daylight
can be supplemented by automatically controlled variable level artificial
lighting.
After a brief description of the principle of lighting control in inductive
circuits, the special demands which must be met by a dimmer for fluorescent
lamps are discussed; these dimmers have to be made a good deal more com-
plicated than those for dimming incandescent lamps.
DIMMING OF FLUORESCENT LAMPS 199
10.2 The principle of dimming fluorescent lamps
As described in the previous chapter, a silicon controlled rectifier (thyris-
tor) is cut off in the reverse direction, just like a normal rectifier, i.e. no cur-
rent of any importance will flow when the cathode is positive with respect to
the anode. Unlike a normal rectifier, however, a thyristor will be cut off in
the forward direction too, until a suitable signal is applied to its gate elec-
trode. The thyristor will then become conducting, and will remain so until
the current falls below a low value (the holding current), when it will become
cut off again.
In order to make the thyristor conducting again, we have to apply another
suitable signal to the gate. Since the silicon controlled rectifier only passes
current in one direction, for lighting control purposes we have to use two
thyristors in anti-parallel- one for each direction of the current (Fig. 10.1).

Fig. 10.1 Principle of dimmer circuit with radio-interference filter. The mains voltage is not
applied across the load, consisting of inductively stabilised lamps, until one of the thyristors
is made conducting by means of an appropriate triggering signal between gate and cathode.
The capacitor in the bridge suppresses mains voltage transients.

The time at which the control signal is applied (or the angle cp in Fig. 10.2)
can be varied by means of a phase shifter. Current can only flow through the
load after this signal has been given.
For a lamp stabilised by a self-inductance we may write:
di
vm sin (wt + cp) = L- + VI
dt
where V 1 is the lamp voltage, which is independent of the current to a first
200 FLUORESCENT LAMPS

Vm sin wt

Fig. 10.2. Dimming is realised each half cycle by applying the mains voltage to the load
at a (variable) time t., a phase angle rp after the mains voltage passes through zero.

approximation. Integration gives:


-Vm
--cos (wt + rp) = Li +V 1 t- C
w
Since i = 0 at t = 0, we then have
vm
C =-cos rp,
w
so that
vm + rp)- rxwt) (10.1)
i =-(cos rp- cos (wt
wL
where
Vz
rx = --
Vrn
The current thus has three components, proportional to cos rp,
-cos (wt + rp) (depends on the switching-on time) and -rxwt (varies
y

a.wt
~~--~~~~~~-=~~-x
Fig. 10.3. The current through a
dimmed lamp, and its components as
given by eq. 10.1, as a function of time.
------------------------cos 'f! The resultant lamp current is given by
the sinusoidal curve, after the whole
graph has been rotated about the origin
so that the line rxwt has become the
abscissa.
DIMMING OF FLUORESCENT LAMPS 201
linearly with time). The overall form of the current is as shown by the
curve forming the upper limit of the shaded area in Fig. 10.3. If the current
passes through zero, the thyristor cuts off automatically, and no current
will pass until the other thyristor is triggered during the next half period.
If the thyristor is triggered later, the current produced can easily be found
from Fig. 10.3 by shifting the line rxwt upwards parallel with itself. The resul-
tant current becomes smaller as t. shifts more to the right, and the lamp
will never ignite if the line rxwt becomes a tangent to the curve-cos (wt + q;).
The time at which the current is maximum can be found by differentiating the
current given by equation 10.1, and equating difdt with 0. This gives:
sin (wt + q;) = rx (10.2)
In other words, the maximum current always occurs at the same time, inde-
pendent of the moment of switching on.
The greatest possible current is produced if the current passes through zero
precisely a half period after switching on, so that there is no currentless time.
The condition i = 0 at t = T/2, substituted in equation 10.1, gives
Vm
- (2 cos q;- nrx) = 0
wL
or
nrx
cos q; =- (10.3)
2
If we take a case likely to occur in practice, with V1 = 110 V and Vm =
220Veff we find rx = VtfVm = 110/220 V2 = 0354. It then follows from
equation 10.2 that the maximum instantaneous current occurs at 159, while
according to equation 10.3 the maximum luminous flux is obtained if the
circuit is switched on at cos q; = 055, i.e. at 57; see Fig. 10.4.
Dimming installations constructed according to the principle described
above give satisfactory results as long as the light is not dimmed too far;
if this is done, mains fluctuations will cause the lamps to flicker, which may
be a serious disadvantage and never looks nice in any case.

Fig. 10.4. The maximum current is obtained when the lamp is switched on at time t 0
(rp = 57). If the lamp is switched on later, e.g. at
t1o the current will be lower and the
lamp will be without current during part of each half cycle. If the thyristor is triggered
at 159, no current at all will flow.
202 FLUORESCENT LAMPS

10.3 Influence of fluctuations in mains voltage


If the angle cp in Fig. 10.3 is gradually increased (more dimming), we can
read off the total conducting time in each half period and the maximum
current from Fig. 10.3, by shifting the line rJ.wt upwards parallel to itself
by the appropriate amount. The product of these two quantities provides
a good measure of the lighting level, and will clearly decrease as cp is in-
creased.
If we assume that the mains voltage substituted in equation 10.1 is e.g.
10% below the nominal value, we get a curve similar to that of Fig. 10.3
for the relation between the lighting level and the angle cp. Comparing the
various curves obtained in this way, we find that the percentage of light at
90% of the nominal mains voltage, with respect to that at the nominal
voltage, varies with cp as shown in Fig. 10.5. It will be seen from this figure
that the light is very sensitive to slight mains variations, especially at low
brightnesses.

% 100

' 90

80

70

60

50

1.0

30

20

10

m
OL-~---L--~--L-~--~---L--J-__L-~--~--~~~--L-

~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ m
--------.. 'fJ
Fig. 10.5. The light output at 90% nominal mains voltage as a percentage of that at
100% nominal mains voltage, as a function of the time during each half period at which
the thyristors are triggered, i.e. as a function of the angle rp of Fig. 10.2.

The reason for this great sensitivity to mains fluctuations lies in the fact
that the lamp, unlike a normal resistance, keeps the lamp voltage practically
constant as the current varies. The situation is illustrated in Fig. 10.6 and
Fig. 10.7. Fig. 10.6 shows the vector diagram of a fluorescent lamp stabilised
by a self-inductance. The mains voltage Vm is equal to the vector sum of
DIMMING OF FLUORESCENT LAMPS 203

Fig. 10.6. Vector diagram for a lamp in Fig. 10.7. The mains voltage and the lamp
series with a self-inductance. The lamp current as functions of time, when the thy-
is regarded for this purpose as a purely ristor is triggered at time t 0 The curves Vm
ohmic resistance, the voltage across and i refer to the nominal mains voltage,
which is kept constant. and the curves Vm' and i" to 90% of the
nominal mains voltage. The curve i' gives
the current at the lower mains voltage if
the thyristor is triggered at the earlier
moment t 1 .

the lamp voltage V1 and the voltage across the self-inductance, V L Because
the operating voltage of a gas discharge lamp is practically constant, the
current is determined by the value of the self-inductance in series with the
lamp, and the voltage across this component: i = VL/wL. When the mains
voltage falls from vm to Vm', the voltage VL will fall to VL' as shown in the
figure. The current will then also fall in the proportion V L, IVL and the
phase angle will also fall, from cp to cp'. This effect is plotted as a function
of time in Fig. 10.7. The control signal is applied at time ! 0 , and causes a
current ito flow. If now Vm falls to Vm', the current will fall from ito i',
the phase angle cp between the current and the mains voltage will fall to cp'
and the curve i' will lie to the left of the curve i. Since however the control
signal is only available at time t = t0 , the actual current which will flow
is i", which is even smaller than i', and much smaller than i.
In view of the above, it is clear that the moment at which the control signal
is applied should if possible be made dependent on the instantaneous value
of the mains voltage. For example, if the mains voltage falls from Vm to Vm'
as shown in Fig. 10. 7, the control signal should automatically be shifted
from time t 0 to t 1 or even earlier, so as to make i' equal to i, i.e. independent
of the mains voltage. A special control circuit for the thyratrons has been
developed to achieve this; we shall discuss this circuit in the following section.

10.4 Control circuit with stabilising effect


In Fig. 10.8, C and R form a phase-shifting network: variation of R causes
the phase of the voltage between A and B to vary with respect to the mains
voltage. A self-inductance and a Zener diode are connected in series between
204 FLUORESCENT LAMPS

Fig. 10.8. Triggering circuit for


the thyristors.

A and B; the latter forms part of a bridge circuit. A Zener diode differs from
a normal diode in that it is cut off in the reverse direction only up to a
certain voltage, above which an appreciable current flows. A Zener diode
thus gives a square wave voltage (like a discharge lamp) when it is connected
to a sinusoidal voltage source in series with a self-inductance. This series
combination can thus be seen as a copy of the fluorescent lamp with its
inductive ballast, and has the same characteristics as the latter. The Zener
voltage is amplified by a transformer with two secondary windings and two
transistors, and is then fed to the gates of the controlled silicon rectifiers.
Now if the supply voltage of the whole circuit fluctuates, the Zener voltage
(like the voltage across the lamps which forms the load) will remain constant,
and the phase angle between Vc-B and VA-B will vary. If the mains voltage
decreases, the point t0 of Fig. 10.7 will shift to the left, and vice versa. When
Vc-BIVA-B = VtfVm (in Fig. 10.8), the point t0 (in Fig. 10.7), correspon-
ding to i, will automatically shift to t 1 while the current changes to i'. When
Vc-BIVA-B is greater than VdVm, to will shift further to the left, i' will
increase and can even equal i if the circuit is designed properly. This means
that the lamp current will be independent of fluctuations in the mains voltage,
within certain limits.

10.5 Symmetry of the lamp current


One of the two thyristors passes current during one half period and the
other during the other half period; it is therefore important that they should
both operate in the same way, i.e. that they should not only start to conduct
at the same moment, but stop conducting at the same moment. If this is not
the case, the current through the lamp will be asymmetrical, causing the
light from the lamp to flicker with a frequency equal to the mains frequency.
DIMMING OF FLUORESCENT LAMPS 205
10.5.1 Start of the conduction
Fig. 10.9 compares the results of controlling a fluorescent lamp and an
incandescent lamp at different moments. After the above discussion, it will
be clear that when the moment at which the thyristor becomes conducting
shifts from t0 to t 1 in Fig. 10.9a, the reduction in the amount of light is
represented by the shaded area.
With an ohmic load (and an incandescent lamp can be regarded for this
purpose as an ohmic resistance), the current is always proportional to the
instantaneous value of the mains voltage (Fig. 10.9b). The shaded area is
much smaller in this case, so that differences between the two half periods
are less important. These considerations thus confirm the above assertion
that special care must be taken to ensure that the two thyratrons used to
control a fluorescent lamp are both triggered at the same moment.

a b
Fig. 10.9. The decrease in the current through a fluorescent lamp (a) and an incandescent
lamp (b) when the thyristor is triggered at t 1 instead of t 0 The shaded area represents the
decrease in the total current.

Fig 10.10. Ideal form of the triggering pulse to ensure


that thyristors with different triggering thresholds
(< V1 ) will all be triggered at the same moment.

Since however different silicon controlled rectifiers require signals of dif-


ferent amplitudes to trigger them, the triggering pulse should have steep
edges as shown in Fig. 10.1 0. With such a pulse, any silicon controlled recti-
fier requiring a triggering signal ~ V1 will ignite at the same time, t 0 This
is automatically realised in the control system of Fig. 10.8 by connecting
the Zener diodes in series with a self-inductance, which gives a current with
a phase lag with respect to the mains voltage. When the current i in Fig. 10.11 a
passes through zero, then, the instantaneous value of the supply voltage will
exceed the Zener voltage, so that the current will be practically uninterrupted
- unlike the case with a Zener diode in series with a resistance, where the
current and the supply voltage pass through zero at the same moment (Fig.
10.11 b). There will be a period without current in the latter case, because the
voltage available at the moment when the current passes through zero is less
206 FLUORESCENT LAMPS

a b
Fig. 10.11. The voltage VA-B between the points A and Bin Fig. 10.8, the voltage v.
across the Zener diodes and the current i through these diodes, as functions of time,
(a) for the circuit of Fig. 10.8
(b) if the choke coil Lis replaced by a resistance.

than the Zener voltage. As a result of this, the voltage pulses produced by
the Zener diode (used for triggering the thyristors) will not have steep flanks.
In order to ensure that the triggering pulses for the thyristors are steep
enough (e.g. a rise time of about 0.1 ms for 8 V), the triggering pulses should
not represent too much of a load for the Zener diodes. In our case, this means
that the current drawn for the triggering pulses should not be more than
5% of the total current. Now the current required for the triggering of a
silicon controlled rectifier is of the order of 100 rnA. If the triggering pulse
for the thyristors came directly from the Zener diodes, the bridge of the phase
shifter would have to handle an excessively high power. This problem is
solved by use of a high power current amplifier, which keeps the edges of
the triggering pulses steep while the power in the bridge remains low. The
principle of the whole circuit is shown in Fig. 10.12.
~------------------~--------------------------~------1---
o~---------------r----------------------~,_------r;r

?t load 0
Fig. 10.12. Complete dimmer circuit for ensuring that the lamp current will be symme-
trical. The phase shifter of Fig. 10.8 may be seen on the left, the two thyristors in anti-
parallel (Fig. 10.1) on the right, and the 2-transistor signal amplifier in the middle.
DIMMING OF FLUORESCENT LAMPS 207
10.5.2 The cut-off at the end of each half period
In most applications, silicon controlled rectifiers are made conducting by
the application of a voltage pulse between gate and cathode. They then
remain conducting as long as they continue to pass current.
When the current through the thyristor falls below the holding current at
the end of each half period, the thyristor becomes cut off. Since however
the magnitude of the holding current varies from thyristor to thyristor, the
current may continue for longer in the one half cycle than in the other, again
causing the light to flicker.
This factor is of particular importance at low brightnesses, where the dif-
ference in current due to the switching off of the thyristors at different times
is no longer negligible compared with the total current. This is illustrated in
Fig. 10.13a, where the shaded area representing the effect of switching off the
thyristor accounts for an appreciable proportion ofthe total current- unlike
the case in Fig. 10.13b, which shows the effect of switching an incandescent
lamp off too soon. When this lamp is dimmed, the filament cools off and
the ohmic resistance drops appreciably. The lamp current thus increases as
the lamp is dimmed, so that the effect of the holding current, if any, is ne-
gligible.
If fluorescent lamps have to be dimmed to low brightnesses, it is a good
idea to maintain the triggering signal for a longer period, so that the thyristor
is not cut off until the supply of charge carriers at the gate ceases (when the
anode current becomes negative). In this way, the lamp current will fall to 0
before the thyristors are cut off. The circuit of Fig. 10.12 is designed so that
the triggering signals are maintained for long enough for this principle to
be realised.

inductive fluorescent load incandescent-ohmic load

a b

Fig. 10.13. The loss of current due to the cut-off of the thyristor when the lamp current
falls below the holding current is of less importance with an incandescent lamp (b) than
with an inductively loaded fluorescent lamp (a).

10.6 Controlling high powers (central control unit)


A fluorescent lamp installation with a total power of up to 2 kW can be
dimmed with one central dimmer (see Fig. 10.14); the variable resistance R
of the above-mentioned phase shifter serves to control the level of the illu-
mination. This resistance can be installed at any convenient spot; and if
208 FLUORESCENT LAMPS

Fig. 10. 14. 2-kW dimmer, regul ated by


means of the variable resistance, which
can be adjusted manually or automatically
as desired.

desired can be remote-controlled by means of servomotors from various


control points.
If however the installation exceeds 2 kW, several dimmers are necessary;
these will be connected to the various phases of the mains supply. The various
resistances can then be mounted on a common shaft, so that the various
groups of lamps can be controlled simultaneously. This method can be used
with e.g. two or three dimmers, but it is difficult to dim a very large number
of lamps uniformly in this way, largely because this requires that the different
potentiometers should be very closely 'synchronised' and that the spread
between the various dimmers should be very slight. For these reasons, a
special central control unit has been designed for the control of high powers.
10.6.1 Operation
The central control unit (Fig. I 0.15) basically consists of one phase shifter
of the type described in the previous section per phase; each phase shifter
again has a self-inductance and a Zener diode in series, but the phase shifting
is .no longer realised by means of a capacitance and a variable resistance,
but by a constant resistance Rand a variable self-inductance (transductor) T 3
This system is somewhat more expensive, but it has been chosen because it
allows the phase shifts on all three phase lines to be exactly synchronised.
A special winding is therefore placed on each of the three transductors, and
these three windings are connected in series. A variable direct current is
passed through this whole circuit.
In each of the three phase shifter circuits, the square wave voltage across
the Zener diode is fed to the two output terminals via an amplifier formed
by S 4 , Tr 1 , Tr 2 and S 5 . The three phases have one common lead, so that
four signal wires come out of the central control unit. A maximum of 50
dimmers can be connected to each pair of leads; in other words, the control
unit can control a total of 150 2 kW dimmers. The individual dimmers
DIMMING OF FLUORESCENT LAMPS 209
R----------------------------------------------------
5--------------------------------------------------
T------------------~-------------------------------------
0--~--------------~-------------------------------------

3 transductor wmdings
in series
Fig. I 0.15. The central control unit for dimming high powers. This comprises a number of
dimmer circuits according to the principle of Fig. 10.14, and a supply unit for the vari-
able d.c._voltage.

now differ from the normal version in that they lack the phase shifter;
however, the amplifier stage and the means for feeding the triggering signals
to the silicon controlled rectifiers are present, of course.

10.7 Outdoor lighting (influence of the temperature)


Until a few years ago, dimmers were mainly used indoors, e.g. in theatres,
hospitals, lecture halls, etc. The ambient temperature in such buildings is
not subject to great variation, so that controlling the lighting level comes
down to controlling the lamp current. At present, however, interest in con-
trolling the lighting level in outdoor installations (e.g. in road tunnels) as a
function of the daylight level is increasing.
The wide temperature variations which can occur under such circumstances
entail the need for special control measures, since the light output now de-
pends not only on the lamp current but also quite strongly on the tempera-
ture; moreover, as the temperature falls the lamp shows an increased tenden-
cy to flicker. Fig. 10.16 shows the variation of the luminous flux (among other
quantities) as a function of the wall temperature of the tube.
Optimum conditions are found with a wall temperature of about 40 C,
which is reached when the lamp burns normally at room temperture (22 C).
210 FLUORESCENT LAMPS

% T20 ...
, , 1,
t 110
,/"
1.--=
~ _,/"
TOO
~ !'-........... w,
I?
)<: ---; v:/~
~
90

80 I ~~ 1'-...rJ> Fig. 10.16. The relative lamp voltage Vz,


1/ ~
~.
current lz, power W 1 and luminous flux
tJ> for a 65 W 'TL' fluorescent lamp with
70 an inductive ballast, as a function of the
I wall temperature of the lamp. The

7
60 mains voltage is kept constant at 220 V.
When the ambient temperature is 25 C,
50 the wall temperature is 425 C. Accord-
ing to international convention, the
17 relative value of all variables is taken
40 as I 00% at this temperature.
TO 20 30 40 50 60 70 80
Lamp wall temperature (C)

If the wall temperature is higher than this (and particularly if it is lower),


the light output is lower.
Since the average temperature out of doors is a good deal less than
20-25 oc in most countries, the first improvement which comes to mind is
the use of a closed fitting, so that the temperature difference Ll T between
the lamp and the outside air is increased. Fig. 10.17 shows by way of example
the influence of a special tunnel fitting which gives a Ll T of about 30 degC
with an average temperature of 10 oc; this brings the operating temperature
of the lamp into the optimum range. Because of its large heat capacity, such

""*- 110
v r-- r-- 25with I
t TOO
/v
./ wind
25withou
wind
90
I
v
80

70
I
I
7
60

50

40
-20 -TO 0 TO 20 30
- Ambient temperature in C

Fig. I 0.17. The luminous efficiency 1J of a tunnel fitting with one 65 W 'TL' lamp, as a
function of the ambient temperature.
DIMMING OF FLUORESCENT LAMPS 211

% 100

t \ 1\
.. TL .. ~6sW

I
/ -
80
I
\ 1/

60 \ I
I~ II/
40 ~ 1/
Fig. 10.18. The minimum permis- \v
sible current (curve I) as a percentage
of the nominal current, the luminous
flux at the nominal current (curve II) 20 v
I 1\
and the minimum lighting level at / '\
full dimming (curve III), all as func-
tions of the wall temperature of the
lamp. Curve III is obtained by mul-
-20
/- v
-10 0
Ill
10 20
~~
30 40 50 60
tiplying the percentages of curves I
and II at each temperature. -oc

a system also gives a certain protection against very rapid temperature


fluctuations.
While the power dissipated by the lamp gives a welcome rise in temperature
in such a closed fitting, things become more difficult when the lamp is
dimmed - and hence, of course, produces less heat.
Fig. 10.18 is instructive in this connection. The abscissa of this graph is
the wall temperature of the tube. Curve II gives the percentage luminous
flux of a 65 W lamp as a function of the wall temperature when the lamp
current has the nominal value. Curve I gives the minimum permissible cur-
rent, as a percentage of the nominal current, if the operation of the lamp
is to remain stable. If the operation were governed by this curve alone, the
operating range at low temperatures would be very limited. Fortunately,
curve II improves the situation: while the minimum current is higher at
lower temperatures, the light output is considerably lower, so that the
lighting level can still be dimmed to quite a low value even though the lamp
current must remain relatively high.
The minimum light output at which the lamp still burns stably can be
read off from curve III; this is obtained by multiplying the percentages of
curves I and II at each temperature.
It will be seen that even at the most unfavourable lamp temperature, the
luminous flux can be dimmed to 9% of the nominal value. With the some-
what shorter 40 W lamp, this minimum value is even lower, namely 4%.
As we have mentioned, curve II gives the maximum luminous flux as a
function of the wall temperature of the tube. If a fitting with a characteristic
like that of Fig .1 0.17 is used, then with the nomimal lamp current the tern-
212 FLUORESCENT LAMPS

perature of the wall of the tube is 30 oc higher than the ambient tempera-
ture. The curve giving the maximum luminous flux as a function of the
ambient temperature can thus be obtained by reducing the temperatures on
the horizontal axis by 30 oc.

10.8 Automatic control


It will be clear that manual control of the lamp current in such outdoor
installations is not advisable, since:
(1) the current at which the light output is minimum varies widely with the
temperature,
(2) over-dimming can cause the lamps to flicker, and
(3) the person in charge of the dimming unit would have to take too many
factors into account.
As a result of this, automatic lighting control tends to be used in outdoor
installations such as road tunnels; the optimum current is then automatically
chosen as a function of the desired lighting level and the prevailing tempera-
ture. The control system should also raise the level of the lighting when the
daylight level outside the tunnel increases, so that the transition from outside
the tunnel to inside (the 'black-hole' effect) is not too extreme. At night,
therefore, the lighting level in the tunnel can be made much lower. The
principle of the automatic control system (for regulating the tunnel lighting
level in accordance with the daylight level) is shown in Fig. 10.19. This repre-
sents a bridge circuit, with two Zener diodes in one arm and two photocells

r r--------------1
f--<f-----------7--L_ J Rf
I
I
I

: cell outside

r------- -----,
' '
:I :R2
icell inside
I

3 tronsductor
windings in series
r
L---~_::-_::-_::-_::-_-_-_-_-_-_-::_-_-_-_-::::_:-~ C> ~
amplifier

Fig. 10.19. Circuit for automatic regulation of the lighting level in a road tunnel as a
function of the light level outside. Each of the cells R 1 and R 2 contains a number of
photosensitive resistors (only one of which is shown in the figure). The cell R 1 is placed
outside the tunnel, and R 2 inside. T 3 indicates the control windings for the three trans-
ductors of a central control unit.
DIMMING OF FLUORESCENT LAMPS 2I3
in the other. One of these photocells is exposed to the daylight, and the other
to the light in the tunnel. The bridge can be adjusted by means of the resis-
tances shown in the same arm as the photocells, so as to vary the ratio of
the two lighting levels and the limiting lighting levels. The object is to keep
the bridge balanced, i.e. to keep the potentials at the points 0 and I as close
to one another as possible. If e.g. the daylight level rises, so that the resis-
tance of the corresponding photocell falls, the potential at I will become
positive with respect to the point 0. This potential difference is fed to an
amplifier, which causes a sharp increase in the current through the trans-
ductors in the central control unit (see also Fig. 10.15). This would lead to
a very considerable increase in the lighting level inside the tunnel, except
that as the tunnel lighting level rises the resistance of the tunnel cell also
falls, and with it the potential at point l. This process is continued until a
new equilibrium position is reached. The potential at l follows that at 0
more closely as the amplification factor of the amplifier is higher. To a first
approximation, therefore, we may state that Rexr = Rint in the new equi-
librium position too.
In order to eliminate the effect of transient changes in the external lighting
level (e.g. the passage of a cloud over the sun), the amplifier is made to
operate with a certain time lag.
The effect of such an automatic control unit is plotted in Fig. 10.20. The
total resistance of the outside cell, together with the series and parallel
resistances shown in Fig. 10.19, is plotted against the illuminance outside
(bottom horizontal axis) as the line R 1 It will be seen that at very low light

Lux inside
TO 2 5 2 TO~
kfl
i
40
38
36
34
32
30
28
26
24
22
20
T8
T6
T4
12
TO L-----------~----------~--~T---~~---r~~
T0 2 2 5 2 5 TO~ 2 5
Lux outside

Fig. I 0.20. Overall resistance of the cell R 1 of Fig. I 0.19 as a function of the illuminance
outside the tunnel, and of R 2 as a function of the illuminance inside the tunnel. Depend-
ing on the illuminance outside, the lighting level inside the tunnel can vary from 20 to
200 lux.
214 FLUORESCENT LAMPS

levels the resistance tends to 36 k.Q, while at very high light levels it tends
to 14 k.Q. The line R 2 represents the resistance of the inside cell together
with its shunt resistance as a function of the lighting level inside (plotted
along the top horizontal axis).
Since as we have seen above R 1 = R 2 at each equilibrium position, it
follows that the lighting level in the tunnel can be regulated from 200 lux
to 20 lux depending on the light level outside, i.e. from the nominal value
to 10% of the nominal value. The lighting level in the tunnel cannot fall
below 10%, since R 1 max = R 2 max = 36 k.Q; this prevents flickering of the
lamps, as remarked in the discussion of Fig. I 0.18 above. It will be clear
that many variants of the circuit discussed here can be devised, depending
on the particular application.
The external photocell together with its associated resistances is placed
in a watertight box. In order to prevent the window of this box from being
covered with snow, which would cause an undesirable reduction in the light-
ing level in the: tunnel, the box is heated to a constant temperature with the
aid of a thermostat.
The resistance combination R 2 is divided between a number of lamp fit-
tings (Fig. 10.22). Rotatable disks with a transmission for visible light which

Fig. I 0.21. The external photocell R 1 of Fig. I 0.19 in its protective housing. In order to
ensure that the control system will not be interferred with by snow on the window above
the photocell, the housing is provided with thermostatically controlled heating. In the
foreground, a photosensitive resistor as used in the cell.
DIMMING OF FLUORESCENT LAMPS 215

Fig. 10.22. The internal


'cell' R 2 of Fig. 10.19 is
distributed over the fit-
tings of a number of fluo-
rescent lamps.

varies tangentially are placed in front of the photosensitive resistors here.


By rotating these disks, we can adjust the relation between the lighting level
and R 2 , and hence between the daylight level and the artificial lighting in the
tunnel.

10.9 Auxiliary equipment


Large installations often require other pieces of equipment apart from the
dimmers discussed above. Since these have nothing to do with lighting con-
trol, strictly speaking, we shall only mention them very briefly here. We have:
10.9.1 The cos qJ switch
In general, the total capacitive load is chosen so as to balance the inductive
load when the lamp is fully loaded. This means that when the lamp is dimmed
216 FLUORESCENT LAMPS

(and the inductive impedance is reduced), the capacitive load will be too
large in proportion. This has no effect on the load in the leads, as the total
current never exceeds the compensated full load current. However, in some
cases the stipulation is made that part of the capacitor should be switched
out of the circuit at one or two settings of the dimmer.
10.9.2 Safety devices
The reliability of the system can be maximised by checking the output
signals from the photocell bridge and from the central control unit. If desired,
a second automatic control system, or manual control, can replace the ori-
ginal system if anything goes wrong.
10.9.3 Twilight switch
A part of the installation can be switched in or out when the daylight
reaches a preset low level.

10.10 Dimmer for low powers


Apart from the dimming equipment discussed above, which meets the
highest requirements as regards dimming range, avoidance of flickering, etc.,
a simpler circuit has been developed for lower powers. This can be used for
the dimming of either 800 W of incandescent lamps or 320 W of fluorescent
lamps; see Fig.l0.23. This differs from the above-mentioned dimming equip-

......
'

Fig. 10.23. Dimmer for up to 320 W of fluorescent lamps or 800 W of incandescent lamps.
DIMMING OF FLUORESCENT LAMPS 217

00----------'

Fig. 10.24. Principle of the control circuit for the 800 VA dimmer.

ment in that smaller thyristors are used, and especially in that the control
circuit is simplified. 'Block regulation' (in which the current pulses through
the gates are of long duration) is here replaced by the use of narrow voltage
peaks. This makes no difference as far as the control of incandescent lamps is
concerned, but with fluorescent lamps it means that a ballast resistor must
be included in the circuit to keep the current through the thyristors sufficient-
ly high, especially at low light levels; otherwise there is a risk that the current
might fall below the holding level, so that the thyristor would be cut off
prematurely. Further the stabilistion circuit discussed in Section 10.4, which
makes the lighting level independent of fluctuations in the mains voltage,
is omitted.

10.10.1 Operation
The principle of the control circuit for the 800 VA dimmer is shown in
Fig. I 0.24. The supply voltage for this circuit is rectified by a bridge circuit.
A d.c. voltage of about 16 V is thus produced across the Zener diode Z, as
shown in Fig. 10.25a. During each half cycle of the mains voltage, the capa-
citor C2 is charged as shown in Fig. 10.25b, until the potential at A exceeds
that at B (see Fig. 10.24); the latter potential is determined by the voltage
divider R 10 -R 11 The transistor Tr 1 then becomes conducting, thus allowing
current to flow through the base- emitter junction of Tr 2 , so that Tr 2 also
becomes conducting in its turn. The capacitor can now be discharged via
R 9 are As a result of the low impedance of R 9 are compared with R 11 and R 10
the potential of the point B falls to a low value. The capacitor can thus dis-
charge quickly to a low voltage (see Fig. 10.25c). However, the potential at
A may not fall below that of B yet ; if it did, the transistors would be cut off
again, the capacitor would be recharged, and so on. A number of pulses
could thus be produced per half period, while the capacitor also runs the risk
of not being completely discharged at the start of the following half period.
This risk is avoided by the presence of the series connection R 8 -R 6 -Trc
TrrR 9 aw through which a (low) current flows, keeping the transistors con-
ducting. The transistors are not cut off until the end of the entire period,
when the Zener voltage falls to zero.
218 FLUORESCENT LAMPS

,
-...\
,
,,
,,
. ..
\ Vm

...,.
,, 0

vz
.

b Fig. 10.25.
(a) The mains voltage Vm and the
voltage Vz across the Zener
diode as functions of time;
(b) The potential at the point A of
f~ : Fig. 10.24. If this exceeds the
~]i potential at point B, the transis-
ll,ilt tor Tr 1 becomes conducting;
c (c) The current peak Ic applied to
the gate of the thyristor when
Tr 1 becomes conducting.

0 2 4 0
Fig. I 0.26. Full circuit diagram of the 800 VA dimmer.
DIMMING OF FLUORESCENT LAMPS 219
A complete dimmer circuit based on the principle of Fig. 10.24 is shown
in Fig. 10.26. The combination CcL 1 is used to suppress radio interference,
while the coupled switches shown at the bottom of the figure allow the
dimmer to be used either for incandescent lamps (GL) or for fluorescent
lamps (TL).

10.11 Dimmer ballasts


In our discussion of the thyristor so far, we have assumed that the only
component connected in series with the lamp is the stabilisation self-induc-
tance. However, every ballast must contain at least one transformer, for
keeping the cathodes at emission temperature. These cathodes are designed
so that the normal (full-power) lamp current is sufficient to keep them at
emission temperature. However, if the lamp current falls, the filaments cool
off too much, so that an extra voltage drop is required to pull the electrons
out of the cathodes. This shortens the life of the lamps (like cold starting)
and also means that the different lamps will have different lamp voltages,
so that the lighting may suffer from a certain inhomogeneity.
The simplest form of ballast is shown in Fig. 10.27.
However, since a transformer is needed anyway, we can make use of it
to drive a starterless circuit. The lamp used should then be either a 'TL'M
type or a siliconised lamp with an earthed auxiliary electrode to aid the start-
ing process.

Fig. 10.27. Dimmer ballast consisting of a normal choke coil and auxiliary transformer.

A ballast of the type shown in Fig. 10.27 is sufficient for one 20 W or


40 W lamp which has to meet normal requirements as regards starting in
not too wide a temperature range. The choke used can be the same as in a
normal starter circuit. When the lamp has to be used under extreme condi-
tions, the transformer will have to be made bigger, as it has to step up the
no load voltage too (Fig. 10.28). The primary of the transformer comes
before the thyristors; this is necessary if the cathodes are to receive a con-
stant heating current. Ballasts for one 40 W or two 20 W lamps are realised
according to this principle. However some other lamp types still give diffi-
culties when they are dimmed to low lighting levels. This is because the
220 FLUORESCENT LAMPS

Fig. 10.28. Dimmer ballast giving a higher ignition voltage.

dimming is produced by triggering the thyristors at a later phase. The


instantaneous mains voltage occurring at this moment must be enough to
re-ignite the lamp. With maximum dimming, the thyristors are not triggered
until about 159 after the mains voltage has passed through zero, as we have
calculated in Section 10.2 and since the thyristor is triggered so late, the
current will have been cut off for most of the preceding half cycle, thus
allowing time for appreciable recombination in the plasma of the gas dis-
charge, which makes the lamp more difficult to re-ignite.
If all the lamps have been in use for some time, reasonable results will be
obtained without special measures having to be taken. In some cases (colder
surroundings, new lamps), and especially with 65 W installations, the lighting
pattern will be irregular at low brightnesses, because some lamps will just
re-ignite at the voltage available, while others will just fail to re-ignite. Under
these circumstances, the circuit of Fig. l 0.29 will ensure a completely uniform
brightness patllern, even at full dimming. The capacitance C 1 in this circuit

Fig. 10.29. Dimmer ballast giving an extra voltage peak to ensure reliable re-ignition each
half period.
DIMMING OF FLUORESCENT LAMPS 221
forms a high frequency tuned circuit together with the portion S 1 of the
self-inductance. As soon as one of the thyristors becomes conducting, this
circuit gives rise to a high frequency peak which is superimposed on the
mains voltage. This ensures that the lamp will ignite immediately.
However, after the thyristor has cut off, C 1 forms a damped oscillating
circuit together with the whole choke and the lamp (which is in the process
of being quenched). The voltage across C 1 is thus reversed, as a result of
which the voltage across the cut off thyristors can become unnecessarily and
undesirably high, depending on the moment of switching off. The combina-
tion R 2 -C 2 therefore ensures that the last mentioned oscillatory circuit will
be over critically damped. When the lamp is burning normally, most of the
mains voltage is applied across C2 , which represents a very high impedance
for the 50 Hz voltage. The losses in R 2 are therefore negligible. However,
C 2 represents a low impedance for the high frequency circuit C 1 -S-lamp,
so that R 2 can now play its role as a damping resistance. Just after the instal-
lation is switched on, when none of the lamps has yet ignited because the
cathodes have not yet reached emission temperature, the only load for the
first few periods is the combination RrCz-C 1 (mainly capacitive). This
circuit is charged to the peak mains voltage in one half cycle and can then
no longer discharge, because the thyristor is cut off. A half cycle later, just
before the second thyristor becomes conducting, the voltage across this
circuit can thus attain twice the open circuit voltage of the ballast. This is
the reason for the inclusion of the resistance R 1 , which can be made quite
large in view of the fact that a half period is available for the discharge.
Chapter 11

Ballast De~sign

Th. Hehenkamp

11.1 Introduction
Although the most important task of the ballast is to stabilise the dis-
charge, there are many additional requirements that may be made. For
example, the ballast must provide the correct conditions for the :;tarting of
tha lamp, with or without the help of a starter. Requirements m:ly, more-
over, be made as to the power factor, current distortion, impedance for
audio frequencies used in mains-borne signalling systems, etc.
To satisfy the often numerous and difficult requirements, many circuits
have been developed, these being dealt with in Chapter 8. Another group
of requirements is constituted by the wishes of the fittings maker and the
user to have available ballasts of small dimensions, low losses, a long life
and very low hum level. It is largely this latter group of requirements which
determines the ballast construction used.

11.2 Iron circuits


Chokes are constructed with iron cores to bring the dimensions down to
within reasonable limits. Since the magnetisation characteristic of the core
BALLAST DESIGN 223
material used is curved, certain differences appear with respect to the con-
ditions used in Section 7.3, namely:
(1) the current distortion increases,
(2) the influence of the mains voltage variations on the lamp wattage is
greater,
(3) the ratio of the short-circuit current to the lamp current is higher.
The extent to which these phenomena appear depends on the magnetic
induction in the core materials and is greater at higher values of induction.
A high value of induction is favourable, however, for smaller choke dimen-
sions (see Section 11.4), so that this value is pushed to the maximum possible
limit.
The influence of an iron core of average induction on the current distortion
is very small as can be seen in Table 11.1. This gives the percentages of the
uneven harmonics produced when a 'TL' 40 W lamp is operated on a refer-
ence ballast and on a good commercial apparatus. Also shown is the result
which would be obtained if an ideal linear self-inductance were to be used.

Table 11.1
measured values with
harmonic values calculated
from eq. (7.9) reference ballast good commercial
ballast
3rd 57% 6 % 75%
5th 20% 15 % 18%
7th 10% 08 % 08%
9th 06% 03 % 03%
lith 04% 025% 03%
13th 03% 025% 02%

The theoretical case is approached closely by both the reference ballast


(see Section 7.5) and the commercial ballast.
As the induction is increased to higher values, the distortion increases
rapidly; Fig. 11.1 shows the third harmonic as a function of the induction.
Also included in the figure is the influence of the induction on the ratio of
short circuit current to lamp current and on the sensitivity of the lamp
wattage to mains voltage variations.
A further consequence of the use of an iron core is the introduction thereby
of certain electrical losses. These can be divided into hysteresis and eddy
current losses. The hysteresis losses depend on the composition and pro-
cessing of the core material; the eddy current losses are moreover effected
by the thickness of the material used to build up the core. In order to mini-
mise the losses pure iron is not used, alloying elements such as silicon being
added. The core is, moreover, divided into thin (05-10 mm) laminations,
separated from each other by a thin layer of insulation. For high apparent
powers a thin paper layer is employed for this purpose; at the apparent
power level required for a fluorescent lamp ballast, however, an oxide layer
on the surface of the lamination is sufficient. It is usual to quote the total
losses of the core material in W/kg at an induction B of 1 or 15 V s/m 2
224 FLUORESCENT LAMPS

Is/I 1 .dWz 3.
2,2 %,---------,-----,---,---,-----.

2,0 501---+

1,8 40

Fig. 11.1. Effect of the amount


of induction in the iron core of a
choke upon ls/1, percentage of
~o OL---_L__ _ J_ __ L_ __L_~o 3rd harmonic and the sensitivity
0,6 0,8 1,0 ~2 1,4 1,6 of the lamp wattage to mains
- B(Vsec m-2) voltage variation.

(10 000 or 15 000 Gauss); the losses at other values of B can be calculated
approximately by assuming them to be proportional to B 2

11.3 Different types of core constructions


There are two principal forms of iron cores in use for chokes and trans-
formers, namely the shell type and the core type. The latter is further sub-
divided into two sorts, with one and with two coils respectively (Fig. 11.2).
It can be shown that, for a given apparent power and with given material
there is practically no difference in the amount of material required and the

Shell type Core type with 2 coils

Fig. 11.2. Conventional choke de-


Core type with one coil
signs.
BALLAST DESIGN 225
losses whether one uses a shell type or a core type with two coils *. A core
type with one coil is always somewhat less favourable in the material required.
The type of core most suitable in a given case is influenced much more by
the further advantages and disadvantages of each type.
In the shell type, the coil is enclosed by the magnetic material more com-
pletely than in the core type, so that the external magnetic field can be much
smaller. The two windows in the lamination represent a greater loss of
material, however, although it is often possible to avoid this by employing
the resulting !-pieces as the yoke (Fig. 11.3) or by stamping E shaped lami-
nations as indicated in Fig. 11.4 to form a double E core.

Fig. 11.3. Example of lamination which involves no loss of material. Two E shaped
laminations are stamped out simultaneously, and the centre strips, shown hatched, then
serve for the yoke.

Fig. 11.4. Example of a lamination which can be stamped with low material losses. The
shaded area represents waste. This arrangement gives E laminations with pre-formed air
gap; the core is here formed from two stacks of E laminations.

11.4 Choke dimensions in relation to losses and insulation tem-


perature
As already mentioned in Section 11.1 it is desirable to combine small
ballast dimensions with low losses and a long life. These three requirements
are mutually contradictory in choke construction, so that the most favourable
compromise has to be sought.

* H. C. HAMAKER and TH. HEHENKAMP, 'Minimum-cost Transformers and Chokes',


Philips Res. Reports 5, 357-394 (1950).
226 FLUORESCENT LAMPS

If one starts from a well constructed choke of given apparent power and
attempts to make this proportionately smaller, it will be found that the losses
increase in inverse proportion to the linear dimensions of the choke *. The
temperature rise of the insulation, however, increases inversely propor-
tionally to even the third power of these dimensions, or, in other words,
inversely proportionally to the weight of the choke.
The voltage over the choke terminals is, according to the well-known
formula:
(11.1)

in which w = 2:n X mains frequency in Hz.


AFe =cross-sectional area of the core in m 2
B =max. induction in the core in V s/m 2
n = number of turns.
The current through the coil is:
J Aeu
IL = - - (11.2)
n
in which J =current density in A/m 2 ,
Aeu =total copper cross-sectional area (of n turns in m 2 ).
If we introduce, further:
a = width of core
h = height of core
b = width of window
c = length of window
FFe = iron space-factor
Feu = copper space-factor
then, from Equations 11.1 and 11.2, the apparent power in volt-amperes
(VA) can be found from the equation:
(11.3)
From Equation 11.3 it will be seen that, for a given volt-ampere value, B, J,
FFe and Feu must be made large in order to keep the core dimensions small.
It will be remembered from Section 11.2 that the influence of the curvature
of the choke characteristic restricts the choice of the induction; the B values
for the commonly employed core materials lie between 09 and 13 V s/m2
The current density J can be increased to the limit imposed by consider-
ations of temperature rise, the space factors FFe and Feu finally, are deter-
mined by the technical possibilities inherent in the design and the production
methods. In practice FFe usually lies between 088 and 095 so that an impro-
vement is hardly possible. The value of Feu is much lower and subject to
more variation. This is dealt with in the next section.
Let us now consider a choke constructed as favourably as possible for an
apparent power VA and the highest possible values of FFe and Feu, i.e. that
the core dimensions a, b, c and h are so determined that a favourable com-
promise has been made between cost-price and electrical losses. It will be

* The following observations are also applicable to transformers.


BALLAST DESIGN 227
found in this case that the iron and copper losses are equal. This choke can
be made smaller most satisfactorily by reducing all four core dimensions
proportionally. The new dimensions become:
a b c h
a' = - ; b' = - ; c' h' =-
p p p p
in which p > 1.
From equation 11.3, to keep VA constant, the product BJ must be changed
as follows:
B'J' =p4 BJ
Since the copper losses are direct proportional to J2, and the iron losses,
in first approximation, direct proportional to B 2 , in order to retain the same
division of the electrical losses it is necessary that:
B' = p 2 B and J' = p 2 J
Bearing in mind the fact that the weights of iron and copper decrease by a
factor p 3 , it follows that the losses W of the choke increase according to
W'=pW
The temperature rise Tis approximately proportional to the quotient of W
divided by the total choke surface area A, thus:
LIT= K W/A
so that, with smaller dimensions, LIT increases according to the equation
LIT'= p 3 LIT.
The consequence of a reduction in choke dimensions is thus an increase
in the losses, and still more important, a rapid increase in the temperature
rise inversely proportional to the weight. If, by the reduction in size, the
induction B' assumes too high a value, a departure must be made from the
favourable distribution of the losses between iron and copper so that Ll T
then increases still more rapidly with increasing values of p.
The increase in electrical losses can often be accepted for the sake of the
advantage in smaller dimensions and decreased weight. The higher tempera-
ture of the insulation is, however, a serious disadvantage, since the life of
the choke is strongly dependent on this. Consequently the reduction of the
dimensions of the choke is usually limited by the wish for a long life. In
Section 11.6 the relation between insulation temperature and life is discussed
in more detail.

11.5 Copper space factor


The copper space factor Feu denotes which portion of the window in the
core is occupied by copper, the other portion of the window being filled with
electrical insulant. Unlike the iron space factor, which is close to unity, the
copper space factor is rather low, usually between 025 and 050. There are
several reasons for this. Firstly, the wire is usually round and this limit the
228 FLUORESCENT LAMPS

space factor to~ = 078 when using layer insulation. The wire insulation
has a thickness of the order of 10% of the wire diameter whereas 15% must
be reckoned for layer insulation. Due to all these causes together the space
factor is lowered to 50 or 60 %. An even lower value applies to the entire
winding because the choke must be insulated properly from the core requi-
ring a coil base inside, paper insulation outside and card board base flanges
or a few mm of air on the sides (Fig. 11.5).
It is obvious that it is important to increase Feu in order to reduce the
size of the ballast. There are three methods to do so: random winding, preci-
sion or orthocyclic winding, and the use of flat rolled wire.

Fig. 11.5. Cross-section of a coil wound with paper insulation between


successive windings. It may be clearly seen that the conductor proper
only fills a small part of the window.

11.5.1 Random winding


In this method the insulation between layers is omitted to gain space. This
is not a logical principle for fluorescent lamp ballasts because they usually
have long, narrow windows and consequently few layers of many turns per
layer. The potential difference between adjacent turns is small and only of
the order of 1 V even at the highest transient voltages. The potential difference
between two layers, on the other hand, may be several hundreds of volts
under those conditions, so the insulation between layers is a necessary addi-
tion to wire insulation unless the wire insulation has been chosen uncom-
monly thick.
This objection is less applicable to ballasts in which a great many layers
are used with relatively few turns per layer, and it is usually in such cases
that random winding is applied.
The gain achieved by omitting layer insulation is relatively small. This is
due to the wire being wound helically on the base. After completing a layer
wound to, say, a counter clockwise helix, the next layer is wound the other
way, i.e. to a clockwise helix. As a result, the wire will show a tendency to
settle between two wires of the underlying layer but is forced to the next
BALLAST DESIGN 229
groove after a short distance. This results in irregularities, which will increase
in every subsequent layer until in the end we can hardly speak of layers. The
final result is a far less compact stacking of the wire than would be expected.

11.5.2 Precision or orthocyclic winding


The draw backs of random winding can be overcome by applying the turns
in a dense cylinder stacking*. To this end the wire should not be wound in
a helix, but in a plane perpendicular to the coil axis. When a turn has been
nearly completed, the wire must be shifted over a small distance (Fig. 11.6).
The space factor will then be increased considerably, not merely because
the layer insulation has been omitted, but particularly because the densest
cylinder stacking entails an improvement by a factor 1.16 relative to the
normal cylinder stacking procedure (Fig. 11.7).

* W. L. L. LENDERS, 'The orthocyclic method of coil winding', Philips Techn. Rev. 23,
365 (1961 /62).
230 FLUORESCENT LAMPS

Q
Fig. 11.7. Cross-section through a coil with normal cylinder stacking (a) and one with
densest cylinder stacking (b) as encountered in orthocyclic winding. If edge effects in the
wire insulation and cross-over are neglected, it may be seen that the space factor is given
by the ratio of the shaded area to the total area of the square ABCD (078) or the equi-
lterial triangle EFG (091).

Yet the ultimate result is somewhat disappointing owing to the constant


thickness of the insulation. If there is a sufficient layer thickness between
two turns of adjacent layers, the same layer thickness will be present between
successive turns within one layer whereas a much smaller thickness would
be adequate. A further drawback of the orthocyclic method is that the coils
must be wound by a precision machine on very accurate jigs and that, before
removing a jig, the coil must be made to cohere, e.g. by baking.

11.5 .3 Winding with flat rolled wire


Yet another method to improve stacking of the wire is the use of wire of
a rectangular cross-section. Such a rectangular section is very hard to realise
and to handle for the diameters that are of importance in ballasts for fluores-
cent lamps. However, rollers can be fitted to the winding-machine which
flatten the round wire just before it reaches the coil (Fig. 11.8). It is a well-
known fact that the wire insulation can be deformed without appreciable
damage * and a normal ballast life can be achieved by using wire that has
been treated in this way.
When this method is applied an insulation can be provided between layers
and the wire insulation can therefore be thin. The gain in space factor relative
to round wire will be about 15 %. Winding is easy because flat wire settles
tidily forming a smooth surface for the next layer so that no special steps
are required to get a uniform coil which is free from cross turns (Fig. 11.9).

* J. HoEKSTRA, 'Properties and applications of enamelled wire', Philips Techn. Rev. 3,


40 (1938).
BALLAST DESIGN 231

Fig. ll.8. Winding with flat rolled wire. The figure shows the wire guide with flattening
rollers, and the coil to be wound. This set-up can only be used for relatively low winding
speeds; at higher winding speeds, the flattening rollers will be driven by the wire.

a
Fig. 1 1.9. Cross-section through a coil wound with fiat rolled wire and insulation between
successive windings. Magnification 30 X. The two light areas under in the figure represent
two core blocks, followed by the cardboard coil former and the successive windings. The
flattened form of the wire may be clearly seen. At the higher magnification (240 X ) shown
in (b), it may be seen that the wire insulation is not damaged by the flattening.
232 FLUORESCENT LAMPS

11.6 Relation between insulation temperature and life


Since 1930 a distinct correlation has been known to exist between the life
of the insulation and the temperature to which it is exposed: in the range of
the working temperature life is approximately halved per 10 ac increase of
temperature *. Later investigations have shown that the correlation of the
two quantities is linear if the logarithm of the life is plotted against the recip-
rocal of the absolute temperature**. This permits a safe extrapolation of
the results of forced life tests. The actual life of an insulation does not merely
depend on the temperature but also on the quality of the insulation, on the
quantity used and on any protective measure taken. As regards quality it
is usual to classify the kinds available. This is an expedient in choosing but
it should be borne in mind that there may be substantial differences within
one class or grade.
The quantity of insulation has much effect on life. The voltages encoun-
tered in ballasts for fluorescent lamps are so low that direct breakdown of
the insulation at holes or cracks is practically impossible even at high tran-
sient voltages. In the end the ballast will become faulty because due to high
temperatures the insulation is broken down chemically and reduced in quan-
tity. In consequence the turns of successive layers can get so close that a
breakdown becomes possible.
The insulation breakdown is usually slowed down considerably if it is
sealed off from the ambient air so that no oxygen can get in. A perfect seal
entails all sorts of problems, but a partial seal such as can be obtained with

2800C
I'-..
240 .......... ..........
T ~""--... .........
......... ..........
f200
760 ""'~"-!!
'""" ""'
........
I'-..
~.......... ..........
720
.........
r"-..""-....,
700 ..........
80

60 I I
7 3 70 30 7 3 70
days 7 3 6 12 - t years
months
Fig. 11.10. Minimum life (3% failure) of impregnated choke coils and
transformers as a function of the insulation temperature; A in air, Bin
a box filled with polyester mixture.

* V. M. MoNTSINGER, 'Loading Transformers by Temperature', Journal AlEE 49, 293


(1930).
** W. BUssiNG, 'Beitriige zum Lebensdauergesetz elektrischer Maschinen', Arch. Elektro-
techn. 36, 333-361 (1942).
T. W. DAKIN, 'Electrical insulation deterioration treated as a chemical rate pheno-
menon', AlEE Trans. 61, 113-122 (1948).
BALLAST DESIGN 233
a moulding resin already gives a substantial improvement (Fig. 11.10) *.
It will be obvious from the preceding lines that the earlier rule of a specific
grade of insulant having a specific maximum operating temperature is incor-
rect. According to modern ideas the insulating system must be assessed by
forced life tests, a better insulation being evaluated by a higher permissible
operating temperature**.

11.7 Cooling of the ballast


After what has been said in the previous sections regarding the influence
of the temperature of the insulation on ballast life, it will be apparent that
the ballast designer must give considerable thought to the question of good
heat dissipation. For a given construction an improvement in cooling can
reduce the temperature rise and thereby extend the life. Alternatively the
improved cooling can be used to permit a decrease in ballast dimensions.
When discussing heat dissipation a distinction must be made between the
dissipation of heat to the surroundings and the internal heat transfer***.
The external dissipation, i.e. from the surface of the ballast, via the fitting,
to the surroundings is determined by the construction of the fitting and the
method of mounting of the ballast in the fitting. The better this transfer the
lower the surface temperature of the ballast. This will be gone into more
fully in Chapter 12.
The internal heat transfer comprises the conduction of the generated heat
from the innermost parts of the choke or transformer to the ballast surface.
Due to the unavoidable resistance of these parts to the transfer of heat, a
temperature difference is set up between the ballast surface and the insulation
of the windings. An increase in thermal conductivity inside the ballast reduces
this difference and is thus a help in reducing the temperature of the inside of
the ballast.
To this end, all air spaces in the ballast can be filled as completely as pos-
sible with a heat conducting impregnating medium and the space between
the choke and its casing with a heat conducting filling compound. A com-
monly employed filling material is a bitumen compound containing an ad-
mixture of a thermally conductive material such as quartz dust, kaolin, etc.
A disadvantage of the bituminous compound, however, is its fairly high
expansion coefficient and its tendency to flow at high temperatures. This
leads to its being forced out of the can on heating, resulting in a less effective
cooling as well as the obvious inconvenience to the user. A filling material
which does not possess these undesirable features is a mixture of polyester
resins with suitable fillers, providing excellent electrical insulation and a high
coefficient of thermal conductivity ****.The material hardens during the rna-
* TH. HEHENKAMP, 'The life of ballasts for gas-discharge lamps 1', Philips Techn. Rev.
20, 59-68 (1958/59).
** VDE 0712, part 2 (April 1967), 'Ballasts for fluorescent lamps', Publication 82 of the
IEC.
*** C. P. HAYES, 'Temperature Problems of Ballasts for Fluorescent Lamps, Lighting and
Lamps', p. 48 (Dec. 1949); TH. HEHENKAMP, 'Warmteprob1emen bij gebruik van fluo-
rescentielampen', Electrotechniek 33, no. 26, 511-530 (Dec. 1955).
****TH. HEHENKAMP, 'Small ballasts for fluorescent lamps', Philips Techn Rev. 18, 279-281
(1956/57).
234 FLUORESCENT LAMPS

Fig. 11.11. Section of a ballast with a filling based on polyester resin.

nufacturing process and cannot soften in use. The risk of leakage is thus absent
and the coils obtain a hard, continuous, protective covering, ensuring a
longer life (Fig. 11.11 ).
The moulding compounds, though enveloping the core and coil, do not
penetrate far into the coil itself, so air inclusions will be left between the turns
and the layers and between the coil and its final insulation and the base, such
air inclusions hampering the dissipation of heat. In bitumen filled ballasts
this effect is usually combated by impregnating the core/coil assembly with
bitumen before assembly, preferably in vacuo. In the case of polyester filled
ballasts there is an excellent solution by introducing the filling compound
by means of a vacuum. The final insulation will then be forced against the
coil and the coil layers will be forced snugly together. This distinctly reduces
thermal resistance.

11.8 Ballasts of small cross-section


Due to the elongated shape of most fluorescent lamps, leading to the use
in many cases of mounting rails and fittings of small cross-section, it is
obviously desirable to have long, slender ballasts. As a first step in this
direction the components of the ballast (transformers, chokes and capacitors)
are mounted in the box lengthwise with respect to each other. Furthermore,
the components themselves can be designed with a small cross-section.
The consequences of a slender design of a capacitor are unimportant, the
volume and amount of material used being only slightly dependent on the
shape chosen.
On the other hand, the most economical shape of transformers and chokes
is approximately cubic. A departure from this shape results in a biggervolume
BALLAST DESIGN 235
and weight. As long as this departure is not too great, the consequences are
relatively small and can be neglected, in comparison with the advantages of
the more elegant shape of the fittings and the associated savings in material.
With a very small cross-section however, the increased amount of material
required for the ballast becomes quite important, whilst other difficulties
appear in the design. The stray magnetic field increases with long, narrow
shapes of ballast, so that the magnetic forces exerted on the box increase.
The elastic and magnetostrictive core movements also increase, so that con-
siderable attention must be paid to the measures to be taken to ensure a
low noise level. These latter difficulties can be avoided by using two small
chokes or transformers in series or parallel instead of one, but this further
increases the amount of material employed and can thus only be used in
extreme cases.
In choosing the cross-section of a ballast it is therefore necessary to weigh
up the advantages in appearance and the savings in material afforded by the
use of fittings of small cross-section against the disadvantages in the design
of the ballast.

11.9 Construction of capacitors


A capacitor consists essentially of two conducting surfaces separated by
an insulating layer. For the voltages and frequencies usually employed for
fluorescent lamp operation the insulating layer consists of an impregnated
paper of special quality, the conducting surfaces being of aluminium foil or,
in some cases, an evaporated aluminium layer on the paper itself.
To ensure, for a given capacity, dimensions as small as practicable, the
thickness of the insulating and conducting layers must be kept small.
The capacity is determined by the formula:
A
C=e-
d
win which A = surface area of conducting layer.
d = thickness of insulating layer.
e = dielectric constant.
The volume Sc taken up by a capacitor of capacity C can be calculated from
this formula and amounts to:
Cd
Sc=-(d+f),
e

in which f is the thickness of the conducting layer. For given values of C


and e, d and f must be made small to ensure a small volume.
The minimum thickness of the conducting layer is determined by the
technical possibilities; the minimum thickness of the paper layer is determined
by the permissible voltage gradient in the dielectric in relation to the desired
236 FLUORESCENT LAMPS

life. The permissible voltage gradient, for a given life and operating tempera-
ture is determined by :
(1) the quality of the paper,
(2) the number of layers of paper,
(3) the method of drying and impregnation of the capacitor,
(4) the impregnating medium.
The most important points to be considered in the choice of paper, are a
low content of conducting particles and low dielectric losses. The conducting
particles represent weak spots in the insulation, which cannot altogether be
avoided. One way of reducing the influence of these particles is to subdivide
the paper layer by using two or more thinner layers with the same total
thickness. Since a thin paper is more difficult to manufacture with a low
content of conducting particles and low dielectric losses, there is a limit to
the utility of this procedure. For the common voltages between 200 and
500 V, the choice of two layers is often the most favourable.
Low dielectric losses are desirable because of the heat generated in the
capacitor as a consequence of these losses, resulting in an increase in opera-
ting temperature. At higher temperatures the losses increase sharply, so that
an unsuitable choice of paper quality and voltage gradient introduces the
possibility that an equilibrium will not be found between the heat developed
and the heat dissipated by the external surface of the capacitor. The tem-
perature and the losses then increase further until the capacitor fails.
The losses in the capacitor are also increased by the presence of air and
moisture in the paper. Initially, because of the fibrous structure of the paper,
these are always present. It is necessary, therefore, to de-gas the capacitor
cells thoroughly in vacuum and then impregnate them with a good quality
oil.
For this purpose both mineral oils and synthetic oils (e.g. chlorinated
diphenyl) are used. The advantage of the synthetic types is a higher dielectric
constant, which is favourable as regards the dimensions and voltage distri-
bution in the dielectric. The disadvantages on the other hand, are a higher
temperature dependence of capacity and higher dielectric losses *.
With correct choice of paper quality and impregnant, and with careful
de-gassing during manufacture, voltage gradients of more than 107 V/m
are permissible. In order to ensure a permanent moisture and gas free con-
dition of the insulation, the capacitor cells are placed in a hermetically sealed
box. The form of this box can, for a given volume, be chosen more or less
as desired; a large surface area is desirable to ensure a satisfactory heat
transfer.

11.10 Capacitor life


The life of a capacitor is not only determined by the quality of the di-
electric material used, but also by the voltage gradient in and the temperature
of the dielectric. A higher voltage gradient decreases the life, as also does, in

* F. M. CLARK, 'Nonflammable Dielectric Organic Compound'. Ind. & Eng. Chem. 29,
698 (1937).
BALLAST DESIGN 237
general, a higher temperature. If, at a given voltage gradient, the ambient
temperature is raised above a certain value (approx. 50 oc depending on
materials and construction) the dielectric losses in the capacitor will rise,
leading to more heat generation. To dissipate this heat the temperature rise
of the capacitor above the ambient will become higher, resulting in a further
increase in losses. After some time a thermal equilibrium is established, the
temperature rise of the capacitor having then reached a value at which the
heat generation and the heat dissipation are equal.
There exists a critical ambient temperature at which a slight increase in
capacitor temperature will result in a rise in heat generation surpassing the
accompanying increase in the heat dissipation possibilities of the capacitor.
There is then no thermal equilibrium possible and the capacitor will quickly
fail. The higher the voltage gradient the lower is the ambient temperature
at which this thermal instability appears (Fig. 11.12). In practice, naturally,
this limit must not be reached, and therefore a safety margin has always to
be maintained. This implies that the capacitor must not be mounted in too
warm surroundings, so that the contribution to the temperature rise of the
capacitor due of the presence of chokes, transformers and lamps must be
kept within strict bounds. For this reason the capacitor is usually mounted
in the ballast at some distance from chokes and transformers so that the
intervening air space impedes the transfer of heat.

140"C
!
r;
1 120 --r-~
TOO ~
'~
80 r---
"'
60
0 10
i
20 30
-E
I
40
"' SOVjp.

Fig. 11.12. The ambient temperature at which thermal instability appears, as a function
of the voltage gradient, for a certain type of capacitor.

At ambient temperatures 20-25 oc below the thermal stability limit, the


life is usually some thousands of hours, whilst a further decrease in tem-
perature results in an increase in life of the order of five times for each
10 ac. At still lower temperatures this effect tends to disappear, so that at
temperatures in the region of 20 ac and lower, a comparatively short life
238 FLUORESCENT LAMPS

can again appear. This is due to the contraction of the impregnant in the
capacitor, resulting in a low internal pressure. At low pressures the chance
of the occurrence of voids is high and the voltage at which a corona discharge
will start in these voids is low. The detrimental effect of a corona-discharge
on the insulation shortens the life considerably. A low internal pressure has
therefore to be avoided, this can be done using a capacitor housing that
permits a relatively large difference in volume at small variations in pressure.
It is customary to determine the life of a capacitor at 30-50% excess voltage
gradient and a temperature of 10-20 oc above the maximum permissible
operating temperature. To obtain a complete picture of the characteristics,
this is insufficient, since the influence of the corona discharge at low tem-
perature is not taken into account. .
A better method is to test the capacitor in the range between the lowest
and highest temperatures, met with in practice, using a temperature cycling
cabinet*.

11.11 Ballast noise


When fluorescent lamps are used in factories, large stores and other places
where a fairly high noise level always obtains, the noise produced by the
ballasts is not of primary importance. In installations in quiet places, such
as hospitals, libraries and living-rooms, on the other hand, it is necessary to
make the ballast construction such that the noise level is extremely low.
The noise is caused by vibration of the iron core and surrounding case
under the influence of the forces exerted on them by the magnetic field.
These forces are proportional to the square of the induction:
K oc B 2 sin 2 wt = t B 2 (1- cos 2 wt)
so that the noise frequency is twice the mains frequency, the instantaneous
value varying between zero and a maximum value proportional to B 2 As a
consequence of the current wave form distortion, frequencies of 300 and 500
Hz (at 50 Hz frequency) and higher, up to several thousands Hz also appear.
The lower frequencies are radiated by the ballast with a very low efficiency,
because their wavelength is large compared with the ballast dimensions.
Moreover, the ear is not very sensitive to these frequencies, so that the noise
impression obtained from a freely suspended ballast is extremely small. When
the ballast is mounted on a large surface (ceiling, fitting), however, the radia-
tion is considerably stronger, particularly when the mounting surface has a
natural resonance frequency, in the neighbourhood of the hum frequencies.
For the higher frequencies the ear sensitivity is much higher and the radia-
tion from the ballast takes place at a higher efficiency than at the lower fre-
quencies, since the wavelengths are of the same order as the ballast dimen-
sions. It is therefore fortunate that the higher frequency components of the
magnetic field are normally of a very small amplitude.

* TH. HEHENKAMP, 'The life of ballasts for Gas Discharge Lamps II', Philips Techn. Rev.
20, 162-169 (1958/59).
BALLAST DESIGN 239

N /----------, S

Fig. I 1.13. Magnetic poles set up in an iron-core cir-


cuit by air-gaps and joints.

If we now turn our attention to the magnetic forces in chokes and trans-
formers, we can distinguish between forces resulting from the main magnetic
field, from leakage fields and from magnetostriction. The main field passes
through the core and produces a magnetic gradient over each magnetic
reluctance. In this way poles are produced which attract each other with a
varying force (Fig. 11.13). Interruptions in the core circuit, particularly,
represent large magnetic reluctances. In addition, the core laminations are
usually divided into two or more parts, in order to make it possible to put
the coil over the core. Small air gaps (butt joints) result which also display
a considerable reluctance.
The leakage field results from the magnetic gradients set up by the main
field over the core components. The most important consequences thereof
are the forces exerted on the box surrounding the ballast, when this is made
of sheet steel. The use of non-magnetic material for the box does not provide
a solution, since the magnetic field then passes through the walls of the
container and exercises forces on the steel components of the fitting.
The magnetisation of the core material produces small length variations
due to magnetostriction which, dependent on the composition of the material
and the induction, can be positive or negative. These length variations are,
for normal core materials and induction values, approximately proportional
to the square of the induction, so that vibrations of twice the fundamental
frequency are again produced.
To keep the acoustic energy delivered by the choke or transformer small,
the forces must be kept as small as possible, the construction must be rigid,
whilst use can also be made of filling materials in the box with high acoustic
damping. For chokes with air gaps there is also a compensation method
possible.
The forces can be kept small by avoiding high magnetic gradients, thus by
avoiding air gaps and joints. The influence of the finite permeability of the
core material can be overcome by winding the coil evenly over the whole
length of the core. This measure also avoids the production of a leakage
field. Unfortunately, such a construction is difficult to realise in practice,
and hence expensive.
Another way to keep the forces small is to choose a low value of core in-
duction. Since the forces are proportional to the square of the induction,
the sound energy is proportional to the fourth power of this quantity. In
order to obtain an audible difference in noise level (3 dB) a decrease of
20 % in the induction is needed, so that the dimensions of the ballast have
to be increased considerably. An appreciable decrease in noise level in this
way is thus also expensive.
High rigidity can be obtained by assembling the laminations asymmetric-
ally, inserting them from different ends of the coil alternately, so that the
240 FLUORESCENT LAMPS
Core type

even numbers uneven numbers


Shell type

Fig. 11.14. Method of overlapping


the joints of an iron core.
even numbers uneven numbers
joints overlap (Fig. 11.14) and the core components must overcome con-
siderable frictional forces in order to move relatively to each other.
A highly damping filling compound between the ballast parts and the box
decreases the transfer of vibration energy to the box walls and also hampers
the box vibrations caused by the stray field. For this purpose bituminous
compound and a filling composition on a polyester basis are extremely
suitable.
One further method, which can be followed when an air gap in the core
is employed, is the use of a compensated construction *. The principle on
which this works is shown schematically in Fig. 11.15.

:I
Fig. 11.15. Compensated choke construction. A suit-
able material placed between the two E shaped sec-
tions of the core forms the 'air gap'. At higher values
of the induction, the 'air gap' is compressed by the
attraction of the two core sections while these sections
extend slightly by magnetostriction. When for all
values of the induction 21' + b' = 2/ + b, perfect

*
:I compensation of the core ends is obtained.

E. W. VAN HEUVEN, 'On the Noise of Fluorescent Lighting Installations'. Acustica, 5,


101-111 (1955).
BALLAST DESIGN 241
The two core sections attract each other with a force proportional to the
square of the induction. They move in such a way, therefore, that the sides
of the air gap approach each other, over a distance which is determined by
the size of this force and the elasticity of the material in the air gap. At the
same time, however, the two core sections extend slightly by magneto-
striction (with the correct choice of core material and induction) to an extent
proportional to the square of the induction. If the various dimensions and
quantities are suitably selected, the ends of the core away from the air gap
will not move and are thus unable to transfer any acoustical energy to their
supports.

11.12 Noise measurement


If we are to measure the noise produced by a ballast, we can either measure
the ballast itself or the assembly of ballast and mounting surface (fitting or
ceiling). The former method has the drawback of the ballast radiating low
frequencies at a very low efficiency so that even in a poor ballast the noise
level can be immeasurably low.
The higher frequencies are emitted at a much higher efficiency and may
therefore be measurable. In practice it is only rarely that the higher frequen-
cies play a part, so that the measurement gives an entirely erroneous picture.
The second method conforms more closely to practical conditions but it
compels us to make another choice: the type of mounting surface to be used.
This choice is important because the mounting surface does not only amplify
the noise because large dimensions may result in an efficient emission of
sound but it may also ~o so to a violent degree as a result of resonance.
This can be explained as follows:
A ballast is a vibrating system which is subject to a mechanical force and
which may be characterised by a mechanical impedance which is a complex
quantity. The real part of this quantity is made up by the damping, the
imaginary part by the mass and the elastic stiffness of the ballast. The
mounting surface is also a vibrating system which can be excited by the
ballast, and which has a mechanical impedance of which the imaginary part
is made up by the mass and the elastic stiffness of the mounting surface.
Coupling of the ballast to the mounting surface will result in an increase of
the amplitudes of the vibrations and thus of the intensity of sound if the
imaginary parts of the impedances of ballasts and mounting surface have
opposite signs, so that the total impedance is smaller than each individual
impedance. This situation is quite common in practice, and is the reason
why it is advisable to secure the ballast to a mounting surface which shows
resonance effects during the measurement of the sound. After these explana-
tions it will be possible to evaluate the various measuring methods used*.
(1) Direct noise measurement, with a microphone and amplifier of a ballast
supported on rubber. To screen off the always present ambient noise,
the ballast can be placed with the microphone in a sound proof box.
A sensitive microphone and considerable amplifications are needed, so
that a careful screening of electrical and magnetic interference fields is
* C. P. HAYER and H. R. GouLD, 'Noise evaluation of fluorescent lamp ballasts, AlEE
Trans., 70, Pt I, 573-579 (1951).
242 FLUORESCENT LAMPS

Fig. 11.16. Simplified circuit dia-


gram of measuring equipment with
vibration pick-ups.
B =ballast
P = vibration pick-ups
M = mixing circuit
D = differentiating circuit

necessary. As already explained, the very importan tlow frequencies figure


much less in the measurements than is the case in practice. This method
is therefore not to be recommended.
(2) Local measurements of the vibration with pick-ups (Fig. 11.16). The styli
of two vibration pick-ups are pressed against the core or the side of the
box of a freely suspended ballast. The signals from the two pick-ups are
mixed, differentiated and amplified, and presented to an oscillograph
and a voltmeter. The measuring equipment is calibrated in microns of
amplitude. It is not easy in this way to obtain a complete picture of the
total noise produced, but for the tracking of points of excessive ampli-
tude and the investigation of the effect on these of measures taken to
decrease the amplitude (such as the compensation method), this type of
measuring equipment can be very useful. In addition, points of small
amplitude can be traced, suitable as points of suspension of the ballast.
For this latter purpose a sensitivity of about 001 [Lm is required.
(3) The measurement of a ballast which has been secured to a suitable reso-
nant substratum of large dimensions. This substratum can be a pipe
made of plywood whose length equals half the wavelength of the funda-
mental in the acoustic spectrum. The microphone can be suitably
accommodated inside the pipe (Fig. 11.17). Another possibility is to fix

Fig. 11.17. Block diagram of equipment for measuring sound by means of a resonator
pipe. B = ballast, M = microphone.
BALLAST DESIGN 243

Fig. 11.18. Sketch of a set-up where the


noise from a ballast mounted on a dummy
fitting is led via a conical pipe of the
measuring microphone.

the ballast to a freely suspended imitation troffer, and to conduct the


sound to the microphone through a tunnel (Fig. 11.18). Both methods
have the drawback of not being readily reproduceable: if the same ballast
is measured several times, the results can vary a good deal. This is due
to variations in the coupling of ballast and mounting surface. By running
a series of measurements with each ballast or with several ballasts of
the same type, a fair evaluation can be obtained.
(4) If- as stated above- it is necessary to measure several ballasts to get
an average of the variation in coupling between ballasts and mounting
surface, it isbettertomeasureagreatnumberofballastssimultaneously. To
this end 8 or 10 ballasts are mounted on large boards of plywood, which
amplify the sound substantially and which have been installed in a very
quiet room (Fig. 11.19). The inner walls of this room should preferably
not be absorbent but should rather have a highly diffuse reflectivity. This
will raise the sound level and distribute it more uniformly throughout
the room so that the sound can theoretically be measured with a few
suitably placed microphones. In spite of the steps taken, the sound level
244 FLUORESCENT LAMPS

Fig. 11.19. Ballasts mounted on large flexible surfaces for listening-tests in a quiet room.

of high quality ballasts is often still too low for a reliable measurement.
In such cases a number of observers can do a listening test. A number
of standard ballasts can then be mounted on the boards for comparison
purposes, but it is a better method to have the sound from standard
ballasts recorded on a tape and introduced into the room through loud-
speakers after amplification. This will permit setting to an identical
impression of sound, provided the acoustic spectra are approximately
the same.
Chapter 12

Installations
J. Funke and J. C. Moerkens

12.1 Temperature problems with ballasts


The performance of fluorescent lamps and their ballasts depends very
much on their operating temperature. Therefore, when mounting lamps and
ballasts or designing fittings, due attention should be paid to the tempera-
ture problem.
How fluorescent lamps are affected by their ambient temperature has been
indicated in Section 2.5, where particularly the influence on the lumen out-
put has been considered. But, dependent on the type of ballast on which the
lamps operate, their arc voltage and current may also vary considerably.
Especially in normal choke operation this is the case, as indicated in Fig.
12.1.
The increase of the lamp current at higher temperatures results in higher
winding temperatures of the choke and this will affect its life, as shown in
Fig. 9.4.
Since the life performance of a choke depends very much on the quality
of the material used (e.g. type of enamelled wire, interleaving paper) no
universal temperature limit can be given. A way out has been found by
relating the permissible maximum operating temperature (tw) of the windings
246 FLUORESCENT LAMPS

100
~
0
~

I
50
40

20 ~
-
.....
10
'~ ~ "'-..f.a=30oC

I ""' ~
';f. 115 I'
7.5

110 5 ~
4 ~ "
3
/.............. .....
1051 t0 =35C
100 2

95
L_~l---rl--~1---+--~1---+1--~

10 20 30 40 50 60C 210 220 230V


- fambient -vm
Fig. 12.1. Increase oflamp current with Fig. 12.2. Minimum lives of ballasts in
rising ambient temperatures (choke a fitting rated at 220 V and an ambient
operation of 40 W lamp). of fa = 30 C. In accordance with
IEC Publication no. 162 (Recommen-
dations for Fittings for Fluorescent
Lamps) this means max. Lit= tw- 30,
measured at 220 V and fa = 30 C.
At is the temperature rise of the
windings.

to the result of an accelerated life test. This procedure is laid down in /EC
Publication 82 (Recommendations for Ballasts for Fluorescent Lamps) where
the concept of tw marking of the ballast is introduced. The idea is, that as
long as the operating temperature of the winding does not exceed tw, a ballast
life of more than 10 years continuous operation may be expected.
The advantage of this method is that high quality ballasts may now be
designed and mounted in such a way that their temperature is much higher
than under the old system where a general limit was fixed. The IEC Publica-
tion specifies the method according to which tw can be determined.
Ballast temperatures are also affected by mains voltage variations. The
combined effect of ambient temperature and mains voltage variations on a
40 W ballast in a typical fitting is illustrated in Fig. 12.2.
In Fig. 12.2 are represented the variations in expected minimum life during
continuous operation of the ballast as a function of the mains voltage and
with the ambient temperature as a parameter. It is assumed that the tw of
the ballast was just reached in the fitting at 220 V mains and at an ambient
temperature of 30 oc. The curves shown relate to permanent operation of
the ballast at the relevant voltages and ambient temperatures. The actual
INSTALLATIONS 247
shape and position of the life lines will depend on the type of ballast and
lamp, as well as on the design of the fitting.
Important factors which will influence the temperature of the ballasts are
the type of fitting (open or enclosed), the number of lamps and the method
of mounting of the fitting. Also the method of mounting the ballast in the
fitting will have a marked influence on the temperature of the windings. As
indicated in Section 11.7 the heat generated in the windings has to be carried
away mainly by conduction and here the existence of a good heat transfer
from ballast to mounting surface should provide a heat sink for the ballast
and the latter should therefore be fixed in good thermal contact over a large
area using all the fixing holes available. How seriously the winding tempera-
ture can be affected by a small air gap between ballast and mounting surface
is illustrated in Fig. 12.3. For thermal reasons it is therefore bad practice
to mount a ballast on rubber washers, as sometimes is done in an effort to
minimise hum.

60

Fig. 12.3. Temperature rise of a typical ballast,


mounted on isolating washers of varying thick-
ness. 05 1.0 l5 2.0mm

12.2 Radio interference


There is hardly ever need to take radio interference into account when
installing fluorescent lamps; interference does occur sometimes, but the fre-
quency with which this phenomenon is observed among the millions of
fluorescent lamps in use is very small. Such interference that does occur is
erratic in nature, and can never be predicted in advance.
Interference with radio receivers is hardly ever a nuisance at frequencies
above 1 400 kHz. It is rarely felt in the short-wave range, and never in
the FM and TV ranges.

12.2.1 Causes of interference

Switching-on effects
As mentioned in Chapter 8, the lamp must receive a sufficiently high igni-
tion voltage together with a sufficient heating current for the electrodes if it
248 FLUORESCENT LAMPS

is to ignite. This is very often realised by means of glow switch starters, which
interrupt the current periodically until the lamp strikes. This can be heard
in a radio receiver, as can the opening and closing of any switch. However,
this interference is not very serious, since it only occurs a couple of times
each time the lamp is switched on.

Interference produced during the operation of the lamp


Here we may distinguish between cathode interference and anode inter-
ference.
The cathodes are designed so as to be optimally heated in normal opera-
tion; they will then produce enough electrons to maintain the passage of
current through the lamp as efficiently as possible. The heat capacity of the
filament serves to keep the filament temperature (and hence the number of
electrons emitted per unit time) fairly constant. However, the electron cur-
rent through the lamp is sinusoidal, which means that hardly any electrons
will be removed from the region of the cathode at the beginning and end
of each half cycle, when the current is low. This leads to the production of
negative space charges in front of the cathode twice a period, giving rise to
a voltage dip and hence to the possibility of oscillation phenomena.
The oscillations which give rise to anode interference occur a few centi-
metres in front of the anode. If the arc voltage is displayed on an oscilloscope,
these oscillations are sometimes visible as slight fluctuations in the horizontal
part of the more or less rectangular wave form. This anode interference is
due to plasma oscillations, and is influenced by the anode design, the rare
gas pressure, impurities in the gas filling and the ambient temperature, among
other factors. These sources of interference can be partly removed by suitable
construction of the lamp; see Chapter 5.

12.2.2 Transmission to the radio receiver*


The high frequency oscillations generated in a fluorescent lamp can be
transmitted in two different ways:

Radiation
The lamp can be regarded as a little transmitter, which 'broadcasts' its
interference direct to the antenna of the receiver. Since however the intensity
falls off with the square of the distance, this means in practice that a distance
of 2-3 m between lamp and antenna is safe. This type of interference can
thus nearly always be eliminated by placing the lamp far enough away from
the antenna. Moreover, receivers with built-in ferrite antennas are not very
sensitive to interference anyway, thanks to the antenna construction.

* See: H. J.. J VAN BooRT, M. KLERK and A. A. KRUITHOF, 'Radio interference from
fluorescent lamps', Philips Techn. Rev. 20, No. 5 (1958/59).
INSTALLATIONS 249
Conduction via the mains leads
The interference signals generated in fluorescent lamps can also be pro-
pagated along the mains leads; in this way, they can influence radio receivers
at much greater distances. This means of transmission is generally the most
important, and has therefore received the most thorough investigation.

12.2.3 Equivalent circuit


The two primary means of suppressing interference are the stabilisation
self-inductance present in every ballast and the capacitor shunted across the
lamp terminals. A number of other factors can influence the interference,
e.g. whether the fitting is earthed or not, the length of the mains leads, etc.
These factors may tend to increase or reduce the interference. A simple high
frequency equivalent circuit can be useful in elucidating the influence of the
shunt capacitor, the choke coil, the wiring and any other measures on the
interference.
We start from the normal electrical circuit of Fig. 12.4a, showing the lamp
in series with a self-inductance connected across the mains terminals. The

ballast

co

"" cs R2

y
cs
0
_g_ _Q. .

~ ~ L
Fig. 12.4. (a) The normal circuit diagram of a fluorescent lamp with ballast and measuring
equipment for investigation of radio interference. This can be reduced, for the purposes
of these investigations, via (b) and (c) to the equivalent circuit of (d). The effect of extra
components added to suppress interference can easily be determined with reference to
(d); in (e) and (f), we see two circuits obtained in this way.
250 FLUORESCENT LAMPS

magnitude of the interference voltage across R 1 and/or R 2 can be measured


via the two decoupling capacitors C4 and C 5 which have a very low impe-
dance for the high frequency signals.
The high frequency equivalent circuit for this set-up is shown in Fig. 12.4b.
We see:
(1) The lamp is replaced by an h.f. generator (interference source) in series
with Z;, representing the internal impedance of the gas discharge.
(2) Co is the interference suppressing capacitor, which is placed in the starter-
switch housing in starter circuits, and in the ballast in starterless circuits.
(3) C 1 represents the h.f. impedance of the coil: successive layers of the coil
form small capacitances in parallel with the coil. The total parasitic
capacitance represents a much lower impedance for the h.f. signals than
the enormous impedance of the self-inductance; the latter may thus be
neglected.
(4) C 2 and C3 represent the capacitances of the beginning and end of the
coil with respect to the housing of the ballast, which may be earthed.
The capacitance of the wiring with respect to earth (which is generally
small) may be included in these quantities.
(5) C6 represents the capacitance of the other mains lead with respect to
earth.

By way of illustration, we give the values of these quantities for a 40 W


lamp on a 220 V/50 Hz mains:
Z; = 300 Q * Co= 104 pF Cz = 160 pF C4 = C 5 = 105 pF
R 1 = R 2 = 150 Q * C 1 = 110 pF c3 = 190 pF
C 5 = 50 pF
The circuit of Fig. 12.4b can be further simplified by the omission of a
number of quantities of minor importance. For example, the impedance of
the two coupling capacitances C4 and C 5 at 1 000 kHz is about 1-! Q, i.e.
much smaller than R 1 and R 2 Further, C3 with a capacitance of 190 pF
has an impedance of about 900 Q at 1 000 kHz, which is large compared
with R 1 ; C 3 can therefore also be neglected. We thus get Fig. 12.4c or, drawn
in a slightly different manner, Fig. 12.4d.

12.2.4 Measures for reducing interference


(1) The starter capacitor forms a voltage divider together with Zi. Increasing
this capacitance reduces the interference voltage across the mains leads.
Calculations have shown that a relatively small capacitance can reduce
the interference in the mains leads very considerably. It has been found
that there is little point in making the capacitance much greater than
10 nF.
The importance of this starter capacitor may be seen from the fact
that about 60% of the (relatively few) complaints of radio interference
investigated were found to be due to defects in this capacitor or to
faulty contacts in the leads to it.

* In accordance with international (C.I.S.P.R.) convention.


INSTALLATIONS 251
(2) The influence of the ballast is reflected by the capacitances C 1 and C 2
Most commercial ballasts are designed so as to meet international speci-
fications in this respect. In special cases, the coils can be made with a
particularly low capacitance. If the coil is split into two adjacent parts,
each part will have a lower capacitance, and since the two (lower) capa-
citances are in series the overall capacitance will be even smaller. Further,
it may be seen from Fig. 12.4c that the current through R 2 is greater than
that through R 1 (IR 2 = lc 1 + fc 2 ); in other words, the interference is
greater in the mains lead without a coil in it. If we split the coil in two
as described above, we get even better results by placing one half in
each mains lead (with the lamp in the middle). However, this solution
is only necessary when very special demands are made on the circuit.
(3) A capacitor shunted directly across the mains terminals (between the
points X and Yin Fig. 12.4a) provides a simple and efficient extra means
of suppressing interference. In the equivalent circuit diagram of Fig.
12.4c, this extra capacitance is connected across R 1 + R 2 , and since it
has a very low impedance for high frequency signals this capacitance
may be regarded as a short circuit, giving the circuit of Fig. 12.4e (from
the simplified circuit of Fig. 12.4d). It may thus be seen that shunting
this capacitor across the mains has the effect on the one hand of causing
R 1 and R 2 to be connected in parallel, so that their total resistance is
only half of that of each one individually; on the other hand, the
influence of the ballast (in particular of C 1 ) is reduced, because the cur-
rent through C 1 no longer flows through R 1 and/or R 2 , so that it can
no longer influence the results of the measurement.
(4) Apart from the capacitor mentioned in (3) above, another capacitor can
be connected between the neutral line and earth. This will be in parallel
with R 2 and C 5 in Fig. 12.4a, i.e. in parallel with R 2 in Fig. 12.4c; see
Fig. 12.4/
As a result, R 1 and R 2 will be practically short circuited, as long as
the auxiliary capacitance is not too small (e.g. 500 nF). This also gives
rise to a very considerable reduction in the interference voltage.
However, the use of this filter is forbidden in most countries, for
reasons of safety (because of the possibility of exchanging the neutral
line with the phase line, and of earthing defects).
(5) Much use is made of the delta filter, which has one capacitor across the
mains (as in (3) above) together with a small capacitor between each
mains lead and earth. These filters are supplied as a single unit. For
reasons of safety, the value of the capacitors connected between the
mains leads and earth is kept quite small (5-10 nF). In the circuit of
Fig. I2.4a, one of the capacitors will be connected between X and Y,
one between X and earth and the other between Y and earth. The h.f.
equivalent circuit of Fig. 12.4/will thus also apply to this circuit; but
instead of one large capacitance across R 2 we have two small ones, one
across R 1 and the other across R 2 Since these capacitances have to be
smaller than that mentioned in (4) above, for reasons of safety, the inter-
ference suppression will also be less.
If a fairly large number of fittings with delta filters are installed in a
long row so that the metal parts of the fittings are connected, all the
252 FLUORESCENT LAMPS

delta filters will be parallel in effect. This is an advantage as far as the


suppression of interference is concerned, but if the row of fittings is
inadequately earthed, it may be dangerous to touch them.
(6) The capacitance between the lamp and the (earthed) fitting has not been
mentioned so far. This capacitance is really distributed along the whole
length of the lamp, but we can assume without making too much error
that the total capacitance is concentrated at the middle of the lamp (half
way along Z 1). In Fig. 12.4c, the influence of this earthing capacitance
can be indicated by connecting one end of a capacitance to the mid-
point of Z 1 and the other end to earth (between R 1 and R 2 ). The h.f.
'generator' can now send interference signals through R 2 via the earthing
capacitance as well as through the above-mentioned capacitances C 1
and C2 The interference level is thus raised.
Increasing the earthing capacitance, e.g. by placing the lamp nearer
its (earthed) fitting, thus raises the interference level somewhat.

12.3 Higher harmonics in the mains current


As has been discussed in Chapter 7, the lamp voltage has roughly a square
wave form. This square wave can be resolved into a fundamental with the
same frequency as the mains together with its odd harmonics as given by
equation 7.1. The resolution of the square wave into its components is
illustrated somewhat more clearly in Fig. 7.3.
The lamp can thus be regarded as a generator producing odd harmonics
of the mains frequency (including the mains frequency itself) in the circuit.
(see Fig. 12.5). The current amplitude of the harmonics falls off sharply with
increasing frequency, because the amplitude of the voltage decreases in in-
verse proportion to the frequency (equation 7.1) and practically all ballasts
contain self-inductances whose impedance ( = 2nvL) increases with the fre-
quency.
Apart from the square wave lamp voltage, the following can be considered
as voltage sources of higher harmonics:
(1) Choke coils which for reasons of economy are designed for operation
far into the saturation region of the magnetisation curve.
(2) Asymmetrical operation of the lamp, which is sometimes encountered.
The lamp voltage then has a somewhat different form in one half period
than in the other, which leads to a second harmonic (generally of low
amplitude) in the current.
(3) Mains voltage distortion, if present. This distortion, which is generally
slight, plays. no role at all in capacitive and inductive circuits, thanks to

ballast

Fig. 12.5. A fluorescent lamp regarded as a generator


of higher harmonics, which gives rise to distortion in
the mains current.
INSTALLATIONS 253
the presence of a self-inductance in these. However, where parallel com-
pensation is used, the capacitor current can be very severely distorted as
a result of mains distortions, because the impedance of a capacitor is
inversely proportional to the frequency.
The IEC has laid down the maximum permissible content of harmonics
in the lamp current shown in Table 12.1; these values apply both to starter
circuits and to starterless circuits, and are based on the assumption that the
mains voltage is sinusoidal.

Table 12.1
maximum value, as a percentage of fundamental, for ballast
harmonic
without (H) marking with (H) * marking

2 5 5
25.
cos 'P 33. cos 'P
3
09
5
7
7
4
~ Q.9
9 3 ~ not limited
11 2
13 and above 1

* The 'H' marking on a ballast indicates that the circuit has a relatively high harmonics
content.

It will be seen that the power factor is involved in the above relation; the
reason for this is obvious. Compensation, e.g. by means of capacitors, invol-
ves compensation of the fundamental only; the higher harmonics remain at
the same absolute amplitude, which means that their relative amplitude
increases.
In order to give an idea of the harmonics contents found in practice,
Table 12.2 gives values measured for the most commonly used types of
ballasts, together with the required values for ballasts without (H) marking.
It will be seen from this table that in general only the 3rd harmonic is of
importance. An exception is formed by those sequence start circuits with a
partial air gap in the iron core (last column of table). As mentioned in
Sections 8.4 and 8.6, in this case the higher harmonics undergo little or no
attenuation from a series self-inductance; but the partial air gap in the iron
core leads to distortion of the open voltage in such a way that the higher
harmonics produced are in antiphase with those in the lamp voltage. In
general, this can be done very successfully for the 3rd harmonics, but not
so well for the still higher harmonics. The spread impedance, which is nearly
always present, must provide the solution here. This ensures adequate sup-
pression of the 9th and 11th harmonics (which were not present to any great
extent anyway), but the 5th harmonic often gives trouble.
If properly designed ballasts are used, the presence of higher harmonics
will have practically no effect, except in the neutral line of three-phase sys-
IV
Vl
~

Table 12.2 Percentage amplitude for different types of ballasts.


Dual lamp Seq. start with
Dual lamp Inductive with Semi-resonance Seq. start
Type Inductive circuit Auto-leak partial air gap
circuit parallel Section 8.6, Section 8.4, Section 8.4,
of ballast ballast Section 8.4, transformer compensation Fig. 8.45 Fig. 8.19
(choke + cap.) Fig. 8.20
Fig. 8.29

Max. Max. Max. Max. Max. Max. Max. Max.


Mea- per- Mea- per- Mea- per- Mea- per- Mea- per- Mea- per- Mea- per- Mea- per-
Harmonic mis- mis- sured mis- sured mis- sured mis- sured mis- sured mis- sured mis-
sured sured
sible sible sible sible sible sible sible sible

2 - 5 09 5 - 5 - 5 - 5 - 5 - 5 - 5
3 75 14 205 27 25 27 10 14 105 25 89 26 215 25 22 27
5 18 7 38 7 09 7 15 7 31 7 14 7 32 7 75 7
7 08 4 04 4 08 4 09 4 14 4 03 4 22 4 29 4
9 03 3 09 3 05 3 05 3 05 3 02 3 08 3 17 3
11 03 2 0-4 2 06 2 04 2 03 2 02 2 03 2 06 2

fl
~
m
"'Q
...,z
~
i!:
'tl
"'
INSTALLATIONS 255
,,,.-- ... ,,
'
''
\
\
\
I \
I \
I \
I
I
I
I
I
I
I
I
I

1 ,
I
I
I
I
I
I
I

Fig. 12.6. The three fundamentals, with phase shifts of 120 with respect to one another,
cancel out in the neutral lead. The third harmonics add up, on the other hand.

terns. Since the three fundamentals are 120 out of phase with one another,
they cancel out. However, the third harmonics add up, as may be seen from
Fig. 12.6. If now the 3rd harmonics make up e.g. 9% of the total current in
the phase leads, the current in the neutral line when the system is symmetric-
ally loaded will be 27% of that in the phase leads, and will consist exclusively
of higher harmonics.

12.4 Lamp performance at low temperature


12.4.1 Ignition
Starter circuits have been discussed in Section 8.4. The ignition properties
of lamps in starterless circuits depend strongly on the design of the ballast
and on the ambient temperature; various cases have been discussed in
Section 8.6.
12.4.2 Operation
Once a lamp has been ignited at low temperatures, and before it has had
time to warm up fully, it may tend to flicker somewhat. This is connected
with poor ballast design. We shall discuss this problem for inductive and
capacitive stabilisation separately; this covers nearly all cases met with in
practice. The possible presence of step-up transformers, resonance circuits,
etc. in the ballast does not affect our considerations.
The relationship between lamp voltage and current has been given in
Fig. 7.1. At normal ambient temperatures, the line V1 = f(i 1) is nearly hori-
zontal, so that to a first approximation we can regard the voltage as constant
as long as the current flows in a given direction. With an a.c. supply, we
then get the well known square wave lamp voltage.
As the ambient temperature falls, the burning voltage increases, and the
curvature at low currents also increases; the peaks in the arc voltage at the
beginning (re-ignition) and end of each half period thus become more pro-
nounced. This has the following consequences:
256 FLUORESCENT LAMPS

Fig. 12.7. If the voltage available at time t 0 is insufficient, the lamp remains cut off until
t 1 , when the mains voltage has risen enough to cause re-ignition.

With inductive stabilisation


If the arc voltage rises at the end of each half cycle of the current, this
means that the phase shift between mains voltage and lamp voltage is
reduced; in other words, when the current passes through zero, the instan-
taneous value of the mains voltage available is reduced (see also Fig. 8.5).
At the same time, the voltage required for re-ignition of the lamp is higher
at lower temperatures. This means that as the ambient temperature is reduced
we reach a point where the lamp no longer re-ignites directly after the lamp
current has passed through zero. This gives rise to two phenomena with
opposite effects: the plasma in the lamp starts to deionise, so that the break-
down voltage rises; and the mains voltage rises to its maximum value.
If the mains voltage rises faster than the breakdown voltage of the lamp,
the latter will ignite, albeit after some delay. This gives the picture shown
in Fig. 12.7. After re-ignition, the lamp voltage falls to its normal value (cf.
the oscillograms of Fig. 8.8).
In the most commonly used circuits for 40 W and 65 W lamps, the above-
mentioned currentless time is encountered when the ambient temperature
falls to about --10 oc. If the peak mains voltage is insufficient to re-ignite
the lamp, it will go out for good; this does not generally occur until the
temperature has fallen well below - 20 C.
In general, therefore, a fluorescent lamp will continue to burn without
flickering, even at low temperatures. Of course, the times without current
in each half period do mean that the power consumption of the lamp wil
fall off.

With capacitive stabilisation


The lamp will re-ignite at the start of each half period if the (falling) voltage
across the self-inductance is high enough (see the text associated with Fig.
INSTALLATIONS 257
8.15). As the lamp voltage increases with falling ambient temperature, the
voltage across the coil will decrease; and when the latter is too low the lamp
will not re-ignite initially. The voltage across the lamp under these conditions
is the sum of the instantaneous value of the mains voltage and the voltage
across the charged capacitor. As the mains voltage vector rotates, its projec-
tion on the time axis changes and the total voltage applied across the lamp
terminals increases. This effect is illustrated in Fig. 12.8 for three different
moments. The maximum value of the voltage in question, Vm + Vc, can
be very high, of the order of 6- 8 times V1 This means that capacitively
stabilised lamps will almost always re-ignite after a certain delay.

1----------
1

l{, r----- , ~t
I \ \
..
''

I
b

Fig. 12.8.
(a) When the lamp current passes through zero in a
capacitive circuit, the voltage V0 = Vm + Vc is
available for re-ignition of the lamp.
(b) If this voltage is insufficient, the lamp remains cut
off. As Vm rotates, v. becomes larger, and may
cause the lamp to re-ignite with some delay.
(c) The maximum re-ignition voltage is achieved when
the vectors V m and Vc are co-linear. If this voltage
is still not enough, the lamp remains extinguished.

Now when the circuit containing a self-inductance, a charged capacitance


and the lamp is closed, the capacitance will initially discharge very fast. The
high current peak passing through the lamp serves to keep the lamp voltage
low during the first half cycle we thus get the same situation as at higher
temperatures (large phase difference, higher coil voltage so that more voltage
is available for re-ignition). The lamp will thus re-ignite normally at the start
of the next half cycle (see Fig. 12.9). The asymmetrical lamp current produc-
ed in this way causes the lamp to flicker unpleasantly. The temperature at
which this occurs depends on the lamp type and the magnitude of the self-
inductance.
Fig. 12.10 gives the coil impedance needed for there-ignition of a capaci-
tively driven 40 W lamp as a function of the bulb wall temperature. The series
258 FLUORESCENT LAMPS

Fig. 12.9. Oscillograms of the lamp cur-


rent and voltage in a fluorescent lamp
which burns asymmetrically as a result
of a delay of ignition in one of the two
half periods.

capacitance is chosen so that the optimum lamp current is obtained at room


temperature, with the coil impedance involved.
It may be seen from Fig. 12.10 that even at room temperature a minimum
impedance of 210 Q is required. This is logical. With a r.m.s. lamp current
of 440 rnA and a mains voltage 10 % below the nominal value, the peak coil
voltage is 09 x 210 x 044 x V2 = 115 V, which is just equal to there-ignition

.Il
~00

wL

360

320

280

2~0

200

160 L - - - - l . - - - - ' - - - - - - ' - -


-20 -10 0 +10 +20 +30

Fig. 12.10. T he minimum coil impedance at which a 40 W lamp burns steadily, as a func-
tion of the ambient temperature.
INSTALLATIONS 259
TLI.OW TLI.OW
% %
130 \l,n=220V 130 \l,n=220V
At 25 C ambient temp.= 100% At 25 C ambient temp.= 100%
120 Leading ballast 120 Lagging ballast

110 ..- ILA


110
------- ...... ./
/"
/"

100 100

90
..
">A
90 11'.A
..,.,~~~~~~.~ VLA
80 80
WLA
70 ip 70

20 30 1.0 50 60 C 20 30 1.0 50 60 C
Fig. 12.11. The lamp voltage, current, power and luminous flux of fluorescent lamps with
inductive and capacitive ballasts, plotted as functions of the wall temperature of the lamp.
All variables are taken as I 00% at a wall temperature of 42 C, which is reached when
the lamp is mounted bare at an ambient temperature of 25 C.

voltage of the lamp. If the coil voltage is lower, the lamp does not re-ignite
immediately at room temperature, and we get the effect of Fig. 12.9.
At lower ambient temperatures, the re-ignition voltage rises; the coil
voltage (and hence the coil impedance) must thus rise with it.
In general, the lamp will remain free from flicker down to about -15 oc
with a standard capacitive ballast.
However, an exception must be made for the sequence start circuits which
make use of distorted transformer voltages to minimise the distortion of the
lamp current (see Fig. 8.20). The absence of a sufficiently large self-inductance
means that these lamps will start flickering when the wall temperature is
about +5 oc. It is therefore inadvisable to use such circuits for street lighting
and similar applications.

12.4.3 Influence of the bulb wall temperature during operation


Since the vapour pressure in the lamp depends on the temperature of the
coldest spot, which will be situated somewhere on the wall of the lamp, it
will be clear that the wall temperature will help to influence the voltage,
current and power consumption of the lamp, and its luminous efficiency.
This is clearly illustrated in Fig. 12.11, which shows these various quantities
as a function of the wall temperature for lamps with inductive and capacitive
ballasts.
It may be remarked that the temperature of the lamp depends on
(1) the power consumption of the lamp. As may be seen from the figure, the
260 FLUORESCENT LAMPS

dissipation in a 40 W lamp accounts for a temperature rise of about


17 degC;
(2) the room temperature;
(3) the operating conditions (type of fitting, number of lamps per fitting, etc.).
The presence or absence of plastic covers on the fittings is of great im-
portance in this connection.

12.5 Dealing with audio frequency signals in the mains


Electricity mains with a frequency of 50 or 60 Hz often carry voltages of
higher frequency, which are used to actuate relays via resonant circuits tuned
to the frequency in question. These relays serve e.g. to switch over from the
day rate to the night rate for particular groups of consumers, fire alarms,
and switching public lighting on and off.
Most of the generators which superimpose these control pulses on the
mains voltage operate at frequencies between 475 and 1 350Hz, but lower
and higher frequencies are also used. Self-inductances do not interfere with
these high frequency signals, as their impedance increases in direct propor-
tion to the frequency. On the other hand, pure capacitances connected to
the mains leads represent a power drain for the audio frequency generators,
and since these generators are generally oflimited power it has proved neces-
sary to lay down minimum impedances at high frequencies for the equipment
which loads the mains. This means that capacitors may not be shunted
across the mains terminals to improve the power factor unless special mea-
sures are taken. Where such shunt capacitors are used, they are connected
in series with reactances; at 50 Hz, the coil impedance is generally so low
that it does not interfere with the operation of the circuit, but at higher
frequencies wL 1/wC, so that the load is substantially inductive. Apart
from shunt capacitors, series capacitors are also used in ballasts. However,
a self-inductance is always included in these circuits to combat distortion
of the current, so that no extra measures are necesary here.
The demands made on circuits as regards these audio frequency signals
vary from country to country. Particularly severe demands are made by the
IEC (the International Electrotechnical Commission), which stipqlates that
the audio frequency impedance of the circuit as a whole must be inductive,
and must have a minimum value of:
(1) for frequencies between 400 and 1 500 Hz: a resistance which would
have the same power consumption as the circuit in question at the
nominal voltage and frequency;
(2) for voltages between 250 and 400Hz: half the value defined in (1) above.
The significance of this can best be illustrated by means of a numerical
example. Let us assume that the power consumption is 100 W, with Vm =
220 VI 50 Hz. This corresponds to two 40 W lamps: together with their bal-
lasts, they will consume a power of about 100 W. The lamp current will be
e.g. 2 x 440 rnA.
The resistance which will have the same power consumption as this com-
bination, as mentioned in 1. above, has the value R = 484 D.
INSTALLATIONS 261
.Il
10'

-+
I 2

I \ I

Fig. 12.12. Impedance as a function of frequency for various circuits. The equivalent
resistance of the lamps is neglected.
Curve I : Only inductive load
VwL = 168 V. l = 880 rnA
Curve 2: Only capacitive load
VwL = 168 V. l = 880 rnA Vcav = 380V V 0 = 75!Hz
Curve 3: Inductive load with parallel compensation
Voux. self- ind. = 14.7 V
fcomp = 076 A
Vcap = 2347 V
C=llO,uF
Cv 0 =200Hz
Curve 4: Minimum IEC requirements.

In Fig. 12.12, line 4 represents the IEC requirements at different frequen-


cies. The impedance of the coil in inductive circuits is plotted as a function
of frequency in curve 1, capacitive circuits give curve 2. To the right of the
resonance frequency of coil and capacitor at 75 Hz, the latter circuits are
inductive, and to the left they are capacitive. Both these curves lie above
the stepped curve representing the IEC requirements, so no further measures
262 FLUORESCENT LAMPS

are necessary. This confirms that all existing ballasts incorporating a series
capacitor satisfy these requirements without special measures being taken,
thanks to the self-inductance connected in series with the capacitor (for other
reasons, see Section 8.4).
The situation is different when parallel compensation is used, as shown in
Fig. 12.13. It is then necessary to include an auxiliary coil Sin the circuit.
In theory, the stringent IEC requirements could be satisfied in this way.
However, calculation shows that the coil required would be so large, and the
capacitor voltage so high, that parallel compensation cannot compete econo-
mically with the use of alternate inductive and capacitive circuits; moreover,
the resonance frequency of S and L must be borne in mind in design con-
siderations.

Fig. 12.13. When a capacitor is shunted


across the mains terminals in inductive
ballasts in order to increase the overall
cos q; an extra self-inductance S some-
times has to be introduced.

The frequency of 75 Hz used in capacitive circuits (curve 2) is suitable


from a technical point of view, but it is too expensive as we have just men-
tioned.
A resonance frequency in the region of 100 or 150 Hz is not advisable, as
these frequencies can be produced as higher harmonics of the mains voltage,
leading to overloading of the compensation circuit. It is therefore advisable
to choose the resonance frequency in the neighbourhood of the fourth har-
monic. With the above-mentioned load of 100 W, this gives curve 3 of
Fig. 12.12; this satisfies most practical requirements, but only meets the
IEC requirements above 1 200 Hz.
INDEX 263

page page
absorption coefficient for UV 42,43 chromatic adaptation of the eye. . 93
- , solid substances . 35 chromaticity diagram . . . . . . 75, 76
-spectrum . . . . . . . 34 C.I.E. 1931 standard system of colori-
a.c.-d.c. converter 181 metry. . . . . . . . 75
a.c. operation . . . . . . 112 C.I.E. test colour method 96
activation of the electrodes 69 circline lamps . . 127
activator 36, 37 climatic test rooms . 196
afterglow 34 coactivator . . . . 39
amalgam 27 coating, 'down flush' 62
-lamps 59 - of silicone 62
anode fall 11 -, optical control 63
apatite . 46 - requirements . 63
applications of transistor inverters 187 -, 'up flush' 62
arc discharge . . . . . 3 - with a phosphor . 62
audio frequency signals . 260 coiled-coil filaments 53
auto leakage transformers 162 cold cathode emission . 3
automatic control 212 cold cathode lamps . 53
auxiliary equipment 215 colorimeter . 77
average lamp 119 colour 71
barium disilicate: lead -inspection 100
47
- strontium disilicate: europium . -map . . . 81
100
--magnesium silicate: lead 47 -points . . 76
- -, computation . 77
- - zinc silicate: lead 47
- zinc silicate: lead 47 - rendering 71, 91
Ba-Sr-Ca-carbonate . - - o f 'standard' fluorescent lamps 98
59
ballast design - temperature (T0 ) 86
222
-life . . . . . . 232, 246 -vision . . . 74
-noise . . . . . . 238 - tolerances 89
conduction band 39
- of small cross-section . . 234
Becquerel . . . . . . . conversion of UV 44
41
black light lamps . . . . . 47, 178 cooling of the ballast 233
Blasse and Bril . . . . . 34 - of the transistor . 186
'cool white' lamp . . 87
blending of fluorescent materials 71
blue halophosphate . . copper space factor . 226, 227
47
blue printing lamps . . 47 corona-discharge . . . 238
binder . . . . . . . 62 cumulative excitation . 6, 11
bituminous compound - ionisation. . . 6, 11
233
Bouma . . . . . . 74,95 current density . . 226
bulb diameter . . . . - distortion . . . 113
53 -form. . . . . 31
-wall temperature 259
- stabilising valve 111
cadmium borate: manganese. 51 current-voltage characteristic . 9
calcium halophosphate 40 'daylight' lamps . . . . . . 86
- orthophosphate: tin 49 d.c. circuits and lamps . . . 172
- silicate: lead 48 d.c.-a.c. inverter . . . . . . 181
- - : lead, manganese . 40, 71, 82
- - : manganese development of de Luxe lamps . . 99
48 dielectric constant 235
- tungstate . . . . . 33, 42, 46, 71, 82
capacitive stabilisation -losses 236
135
dimmer(s) . . . . 216
capacitor . 115, 235
-life - ballasts . . . . 219
236 dimming . . . . 198
caps . . . 60
Casciarolo . 41 direct-current supply I 09
cathode current dispersion . . . . 62
152 distortion factor . 108
-fall . 4, 10
- transformer . 155 -of fundamental 140
dual-lamp circuit . 141
cathodoluminescence 33
chemical composition of phosphors . 41 earth-alkaline carbonates 69
chemiluminescence . . . . . . . . 33 eddy current . . . . . . 223
264 FLUORESCENT LAMPS

page page
efficiency . . . . 25, 109, 110 hot cathode . . . . . . . . . 53
elastic collisions 7 --lamps, starting voltage of . 16
electrodes . . 53 hysteresis 223
- dimensions . 54
-mounts . . . 61 ignition . 14
electron emission . 1, 53 incandescence 32
-hole . . . . . 39 irl.ductive stabilisation . 128
electron-volt . . . 4 inelastic collisions 7
electronic ignitor . 164 instant start circuits 168
emission current . 54 --lamps, starting voltage of 16
-curve 54 ionisation . . . 5
-, solid substances . 35 iron circuits . . . . 222
- temperature . 54 iron space-factor . . 226
end blackening . . . 59 isotemperature lines 90
energy balance . . . 23
- level diagram of Hg 6 lamp design . . . . 52
equal energy spectrum 79 - dimensions . . . 20
erythemal lamps . . . 47 - for special purposes 177
ethylcellulose 62 - for starterless circuits 146
europium activated phosphors 49 - for switch start operation . 125
excitation energy . . . . . 40 -life . . . . . . . . . . 116
-spectrum . . . . . . . 34 -manufacture . . . . . . 52, 61
exhaust machine processing 67 - performance at low temperature . 255
extremely low temperatures 166 lead-activated silicates 47
leakage field . . 239
flat rolled wire . 228, 230 - transformer . 133
flicker 112, 118 lehr . . . . . 61
'floating' ring . 58 Lenard . . . . 41
fluorescence . . 32 low-resistance electrodes . 55, 156
fluorescent dyes 100 luminescence 32
fluorite . . . . 42 luminous ripple 145, 181
forced commutation 192
form factor . 108 Macadam . . . 96
Fourier series 105 magnesium arsenate 100
fundamental . 114 -arsenate: manganese 48,49
- germanate: manganese 48, 99
gaseous discharges 1 - tungstate . . . . 43, 46,47
gas losses . . . . 7 magnetisation . . . . . 222
general colour rendering index 96 magnetic induction . . . 223
geometry of the electrode 57 magnetostriction . . . . 239
germicidal lamps . 178 matching leak transformer . 132
gettering action 62 manganese 37, 71
glow curve 35, 39 mercury . . 59
- switch starter . 123 -flush . . 68
graphic arts industry 103 -spectrum 82
Grassmann 74 - vapour pressure, optimum 26
Guild . . . . . 75 meta-stable levels . . . . . . 7
minimum perceptible colour differ-
halophosphates 43,46 ence . . . . . . . 89
harmonics . . 113, 115, 140, 223, 252 mixing coloured lights 80
-contents . 253 museum lighting . . . 103
heat transfer . . . . . . . 233
Helmholtz . . . . . . . . 74 'natural' lamp colour . 87
higher frequencies 114, 115, 170, 181 NBS unit of colour difference 96
highly loaded lamps 26 negative voltage-current charac-
high resistance electrodes . . . 56, 157 teristic . . 104
horizontal exhaust machines . 68 neon signs . . . . 45
hospital lighting 103 nitrocellulose 62
host crystal . . . . . . . . 35, 37 noise measurement 241
INDEX 265

page page
open-circuit voltage. . . . 152 series inverter . . . . . 194
optimum coating thickness 66 short-circuit current 110
orthocyclic 228 silicon controlled rectifier 199
-winding 229 sky signs . . . 33
oscillograph tubes 45 'slimline' lamps 149
over-lehring . . . 67 space charge . . 8
specification of the colour rendering
parallel compensation . 133 properties . . . . . 95
-inverter . . . . . 194 spectral band method . 95
particle size, phosphor 65 spectrum locus . . . 78
Paschen curve 17 - , solid substances . 35
peak factor . . . . 116 square wave . . . . 105
Penning effect . . . 18 stabilisation . . . . . 11, 104
persistence of colours 93 'standard' fluorescent lamps 88, 98
phosphors . . . . . 33 starter circuits . . 128
- , absorption coefficient 65 -switches . . . 121
- , scattering coefficient . 65 starterless circuits 152
phosphorescence . . . 32, 35, 39 starting methods . 14
photoconductivity . . . 36 step-up transformer . 131
photo copying lamps . . 179 Stokes' law . . . . 36, 37
photoluminescence . . . 33 stray magnetic field . . 235
photosensivitve receptors 74 stroboscopic effects . 118, 196
Planck . . . . 85 strontium magnesium orthophos-
polyester resins 233 phate: tin 49
positive column . 8, 11 - orthophosphate: europium 49
potassium chloride: thallium 38 - pyrophosphate: europium 49
potential gradient 8 structure of the atom 4
power factor . . . . . . 108, 114 suspension 61
preparation of phosphors 49 switching losses 184
processing in the lehr 66 -speed 185
switch start lamps 125
quantum efficiency . . 19, 23, 45
- - of phosphors . 44 television picture tubes 38
quenching . . . . 35
temperature problems. 245
radio interference 247 -rise . 226
random winding . 228 template 77
rare gas . . . . . 30, 59 thallium. 38
reconstituted daylights 96,97 thermal conductivity 233
recovery current . . . 191 -emission . . . 3, 54
reference ballast . . . 119, 223 - instability. . . . 237
reflection of UV . . . 44 - starter switch . . 124
- factor for visible radiation 91 thermionic emission 10
reflectorised lamps . 62 - work function . 2
resonance radiation . 20, 36 thyristor . . . . . 187
resistors. . . . 109, 112 transistor inverter 182
rotational levels 35 triple coil electrode . 58
- - filaments . 53
safety starters . 123 turn off time . 190
scattering of UV 43 turn-on speed . 190
scheelite . . . . 42
self-commutating inverters. 193 ultraviolet radiators 53
self-inductance . . . . 114 under-lehring . . . 62
semiconductors 182 uranium compounds 74
semi-resonance circuit . 157 U-shaped lamps . . 127
sensitiser . . . . . 40 valence band 39
sensitised phosphors . 40, 46 vertical exhaust machines 68
sequence-start circuit . 137, 163, 169 vibration . . . . . 238
series and parallel heating of the elec- vibrational levels . . 35
trodes . . . . . . . . . . . . 153 visibility curve V(Jc) . 75, 78
266 FLUORESCENT LAMPS

page page
visual efficiency . . . . . . 22 Young . . . . . . . . 74
voltage-current characteristic, nega- yttrium oxide: europium 49
tive. . . . 104 - vanadate: europium 49
volume losses 7

wall losses 8 zinc beryllium silicates . . 45


'warm white' lamps . 87 - blendes . . . . . . . . 42
'watch dog' starter 123 - cadmium sulphide: silver 38
'white' light 84 - silicate . . . . . . 38, 42, 45, 71
Wien . . 85 - silicate: manganese 38
willemite 33, 42, 45, 82 - sulphides . . . 33, 42
Wright . 75 -sulphide: silver . . 38, 42

You might also like