SUMMARIZING Informatica

You might also like

Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 140

Prof.

Elsa Rangel de Zambrano


 Palabras señal
 Identificación del Tema
 Ideas centrales
 La estructura del párrafo
 Esquema básico
 Lectura de gráficos

San Cristóbal, Agosto de 2006


Son elementos de transición que nos indican la dirección que toman las ideas del autor
en un párrafo.
Veamos cuales son y para que se usan:

Para mostrar información específica Para mostrar ideas generales


o de descripción
A few of these are Basically Thus
Especially Generally speaking Consequently
To illustrate
For example In general
For instance In conclusion
In particular On the whole
An example In short
In summary
Therefore
Para mostrar organización espacial
Prepositions:
Above Bottom
Across Center
Against Corner
Along Diameter
Around East
Beneath Edge
Beside Exterior
In Facing
In front (of) Front
In the center Horizontal
Inside Interior
Near North
On Parallel
On either side (of) Perimeter
On top of Section
Outside Side
Over South
Surrounded by Symmetrical
Under Top
Nouns and adjectives: Vertical
Back West
Base
Para mostrar contraste:

Diferencias: Yet
But Conversely
However Instead / instead of
On the other hand Bigger than
Difference Although
Less ……than Semejanzas:
Unlike Like
Whereas As in
While In the same way
In contrast Similarly/ Similarities
Differs from… Also
Is different from /than Both
Rather Alike

Para mostrar Orden Cronológico:

By ( +specific time) After


During (+period of time) Before
From (+ specific time) until (+ specific eventually
time) First
In ( +year, month , season, decade) Last
On (+day) Soon
Afterwards While
Later At last
Next Finally
Subsequently Times
Then Dates

Para mostrar clasificación: Para mostrar proceso:

Categories First, second, third


Classes To begin
Groups After that
Kinds Afterwards
Types At that time
Later
Next
Subsequently
Then
Para mostrar Causa and Efecto
Para presentar causas:
Because Is caused by
If Because of
Since Due to
As a result of Is the cause of

Para presentar efectos:

As a result Leads to
Consequently Cause
For this reason Gives rise to
So Brings about
Therefore Results in
Thus Produces.
Causes
Para concluir
 To sum up,
 In summary,
 To conclude,
 In closing,
 Finally, it may be concluded…
 To summarize,
 All in all,
 Overall, it may be said…
 Taking everything into account,
 On the whole,
 All things considered,
 In sum,
 Everything considered,
 By and large,
 In the end,
 In conclusión
 Finally

Ejemplo:

Recently, there have been some significant advances in the fabrication and
demonstration of individual molecular electronic wires and diode switches. This
paper shows how these demonstrated molecular devices might be combined to
design molecular –scale electronic digital computer logic.
Tópico de la Lectura
Cuando se lee es importante saber acerca de que se está leyendo a fin de poder
entender rápida y más eficientemente el material objeto de la lectura.
Cuando se establece el tópico, es decir el tema tratado se conoce la mitad de lo que
se necesita conocer, el siguiente paso es establecer lo que piensa el autor sobre el
tópico y entonces se tendrá la idea principal del tema.
De este modo se puede saber que es importante y que puede desecharse y aún
hasta hacer el resumen del texto. Establecer el tópico de cada párrafo en una lectura
es útil para así poder llegar al tópico general.
Se debe dar el tópico de la forma más exacta posible tratando de no ser muy
específico o muy general en cuanto al tema.
Las Oraciones tópicos
La topic sentence es una oración que contiene la idea principal de un párrafo. Es
con frecuencia la primera oración ; especialmente cuando se argumenta un punto que
puede muy bien ser seguido por más información , ejemplos, etc. Si la escritura esta
explorando un punto , este frecuentemente viene como última oración mostrando una
conclusión desde el argumento.
Topic sentences:
1. Introducen el tema del párrafo y lo limitan para que pueda ser escrito en un
solo párrafo .
2. Son las más generales y más importantes oraciones .
3. Contienen ideas control ( ideas que serán explicadas , definidas, clarificadas o
ilustradas en las oraciones que siguen a la topic sentence)
4. La mejores topic sentences contiene una opinión “probable” usando hechos de
fuentes confiables .
5. También comienza con un propósito.
Cuando tu lees, tratas de conectar lo que estas leyendo con lo que ya conoces .En
relación a hacer esas conexiones rápidamente y con exactitud. Es importante saber
lo que estas leyendo por ejemplo:
e.g.
ENVIROMENT Tema general
Pollution Tema Específico
Ocean pollution Tema específico
Oil spills Sub-tema 1
1989 Alaskan oil spill. Sub-tema 2
Ejercicio de ejemplo:
_____________________________________________

The electrons of different types of atoms have different degrees of freedom to move around. With some
types of materials, such as metals, the outermost electrons in the atoms are so loosely bound that they
chaotically move in the space between the atoms of that material by nothing more than the influence of
room-temperature heat energy. Because these virtually unbound electrons are free to leave their
respective atoms and float around in the space between adjacent atoms, they are often called free
electrons.

In other types of materials such as glass, the atoms' electrons have very little freedom to move around.
While external forces such as physical rubbing can force some of these electrons to leave their respective
atoms and transfer to the atoms of another material, they do not move between atoms within that material
very easily.

This relative mobility of electrons within a material is known as electric conductivity. Conductivity is
determined by the types of atoms in a material (the number of protons in each atom's nucleus, determining
its chemical identity) and how the atoms are linked together with one another. Materials with high electron
mobility (many free electrons) are called conductors, while materials with low electron mobility (few or no
free electrons) are called insulators.
The expression of numerical quantities is something we tend to take for granted. This
is both a good and a bad thing in the study of electronics. It is good, in that we're
accustomed to the use and manipulation of numbers for the many calculations used in
analyzing electronic circuits. On the other hand, the particular system of notation
we've been taught from grade school onward is not the system used internally in
modern electronic computing devices, and learning any different system of notation
requires some re-examination of deeply ingrained assumptions.

First, we have to distinguish the difference between numbers and the symbols we use to
represent numbers. A number is a mathematical quantity, usually correlated in electronics to
a physical quantity such as voltage, current, or resistance. There are many different types of
numbers. Here are just a few types, for example:

WHOLE NUMBERS:
1, 2, 3, 4, 5, 6, 7, 8, 9 . . .

INTEGERS:
-4, -3, -2, -1, 0, 1, 2, 3, 4 . . .

IRRATIONAL NUMBERS:
π (approx. 3.1415927), e (approx. 2.718281828),
square root of any prime

REAL NUMBERS:
(All one-dimensional numerical values, negative and positive,
including zero, whole, integer, and irrational numbers)

COMPLEX NUMBERS:
3 - j4 , 34.5 ∠ 20o

Different types of numbers find different application in the physical world. Whole numbers
work well for counting discrete objects, such as the number of resistors in a circuit. Integers
are needed when negative equivalents of whole numbers are required. Irrational numbers
are numbers that cannot be exactly expressed as the ratio of two integers, and the ratio of a
perfect circle's circumference to its diameter (π) is a good physical example of this. The
non-integer quantities of voltage, current, and resistance that we're used to dealing with in
DC circuits can be expressed as real numbers, in either fractional or decimal form. For AC
circuit analysis, however, real numbers fail to capture the dual essence of magnitude and
phase angle, and so we turn to the use of complex numbers in either rectangular or polar
form.

If we are to use numbers to understand processes in the physical world, make scientific
predictions, or balance our checkbooks, we must have a way of symbolically denoting
them. In other words, we may know how much money we have in our checking account,
but to keep record of it we need to have some system worked out to symbolize that quantity
on paper, or in some other kind of form for record-keeping and tracking. There are two
basic ways we can do this: analog and digital. With analog representation, the quantity is
symbolized in a way that is infinitely divisible. With digital representation, the quantity is
symbolized in a way that is discretely packaged.

You're probably already familiar with an analog representation of money, and didn't realize
it for what it was. Have you ever seen a fund-raising poster made with a picture of a
thermometer on it, where the height of the red column indicated the amount of money
collected for the cause? The more money collected, the taller the column of red ink on the
poster.
This is an example of an analog representation of a number. There is no real limit to how
finely divided the height of that column can be made to symbolize the amount of money in
the account. Changing the height of that column is something that can be done without
changing the essential nature of what it is. Length is a physical quantity that can be divided
as small as you would like, with no practical limit. The slide rule is a mechanical device
that uses the very same physical quantity -- length -- to represent numbers, and to help
perform arithmetical operations with two or more numbers at a time. It, too, is an analog
device.

On the other hand, a digital representation of that same monetary figure, written with
standard symbols (sometimes called ciphers), looks like this:

$35,955.38

Unlike the "thermometer" poster with its red column, those symbolic characters above
cannot be finely divided: that particular combination of ciphers stand for one quantity and
one quantity only. If more money is added to the account (+ $40.12), different symbols
must be used to represent the new balance ($35,995.50), or at least the same symbols
arranged in different patterns. This is an example of digital representation. The counterpart
to the slide rule (analog) is also a digital device: the abacus, with beads that are moved back
and forth on rods to symbolize numerical quantities:
Lets contrast these two methods of numerical representation:

ANALOG DIGITAL
------------------------------------------------------------------
Intuitively understood ----------- Requires training to interpret
Infinitely divisible -------------- Discrete
Prone to errors of precision ------ Absolute precision

Interpretation of numerical symbols is something we tend to take for granted, because it has
been taught to us for many years. However, if you were to try to communicate a quantity of
something to a person ignorant of decimal numerals, that person could still understand the
simple thermometer chart!
The infinitely divisible vs. discrete and precision comparisons are really flip-sides of the
same coin. The fact that digital representation is composed of individual, discrete symbols
(decimal digits and abacus beads) necessarily means that it will be able to symbolize
quantities in precise steps. On the other hand, an analog representation (such as a slide
rule's length) is not composed of individual steps, but rather a continuous range of motion.
The ability for a slide rule to characterize a numerical quantity to infinite resolution is a
trade-off for imprecision. If a slide rule is bumped, an error will be introduced into the
representation of the number that was "entered" into it. However, an abacus must be
bumped much harder before its beads are completely dislodged from their places (sufficient
to represent a different number).

Please don't misunderstand this difference in precision by thinking that digital


representation is necessarily more accurate than analog. Just because a clock is digital
doesn't mean that it will always read time more accurately than an analog clock, it just
means that the interpretation of its display is less ambiguous.

Divisibility of analog versus digital representation can be further illuminated by talking


about the representation of irrational numbers. Numbers such as π are called irrational,
because they cannot be exactly expressed as the fraction of integers, or whole numbers.
Although you might have learned in the past that the fraction 22/7 can be used for π in
calculations, this is just an approximation. The actual number "pi" cannot be exactly
expressed by any finite, or limited, number of decimal places. The digits of π go on forever:

3.1415926535897932384 . . . . .

It is possible, at least theoretically, to set a slide rule (or even a thermometer column) so as
to perfectly represent the number π, because analog symbols have no minimum limit to the
degree that they can be increased or decreased. If my slide rule shows a figure of 3.141593
instead of 3.141592654, I can bump the slide just a bit more (or less) to get it closer yet.
However, with digital representation, such as with an abacus, I would need additional rods
(place holders, or digits) to represent π to further degrees of precision. An abacus with 10
rods simply cannot represent any more than 10 digits worth of the number π, no matter how
I set the beads. To perfectly represent π, an abacus would have to have an infinite number
of beads and rods! The trade off, of course, is the practical limitation to adjusting, and
reading, analog symbols. Practically speaking, one cannot read a slide rule's scale to the
10th digit of precision, because the marks on the scale are too coarse and human vision is
too limited. An abacus, on the other hand, can be set and read with no interpretational errors
at all.

Furthermore, analog symbols require some kind of standard by which they can be
compared for precise interpretation. Slide rules have markings printed along the length of
the slides to translate length into standard quantities. Even the thermometer chart has
numerals written along its height to show how much money (in dollars) the red column
represents for any given amount of height. Imagine if we all tried to communicate simple
numbers to each other by spacing our hands apart varying distances. The number 1 might
be signified by holding our hands 1 inch apart, the number 2 with 2 inches, and so on. If
someone held their hands 17 inches apart to represent the number 17, would everyone
around them be able to immediately and accurately interpret that distance as 17? Probably
not. Some would guess short (15 or 16) and some would guess long (18 or 19). Of course,
fishermen who brag about their catches don't mind overestimations in quantity!

Perhaps this is why people have generally settled upon digital symbols for representing
numbers, especially whole numbers and integers, which find the most application in
everyday life. Using the fingers on our hands, we have a ready means of symbolizing
integers from 0 to 10. We can make hash marks on paper, wood, or stone to represent the
same quantities quite easily:

For large numbers, though, the "hash mark" numeration system is too inefficient.
INSTITUTO UNIVERSITARIO DE TECNOLOGÍA
AGROINDUSTRIAL

NUMBERS AND SYMBOLS

SIGNAL WORDS

Prof. Elsa Rangel de Zambrano

SAN CRISTÒBAL, ______, _______, ______


ACTIVITY I (4 points)

1- Determina en el texto dado las palabras señal que encuentres en el texto y


clasifícalas dentro de cada uno de los grupos dados en el grammar 1.
Signal Word classification meaning
1-
2-
3-
4-
5-
6-
7-
8-
9-
10-
11-
12-
13-

Lee rápidamente el texto y determina lo siguiente:


1-De que trata el texto?__________________________________________________
2- Determine las ideas desarrolladas por el autor sobre el tema.
3-Traduzca el texto.
4-Responda las siguientes preguntas
a)Why is a good or bad thing to talk about expression numerical quantities in
electronic?_______________________________________________________________
________________________________________________________________________
________________________________________________________________________
________________________________________________________________________
b) How many kinds of numbers are mentioned by the author ?____________________
________________________________________________________________________
________________________________________________________________________
c)What are the examples of analog representation and a digital representation of a
number in the reading?_____________________________________________________
________________________________________________________________________
d)What do the analog symbols require? _______________________________________
________________________________________________________________________
e) How can you illustrate the divisibility of analog versus digital representation?
_______________________________________________________________________

NAME:_____________________________ Score:______________________________
_____________________________
IDENTIFICACIÓN DE LA IDEA PRINCIPAL
La comprensión de la estructura de un párrafo y la habilidad de hallar su idea
principal contribuyen mucho a que el lector obtenga mayor provecho de la
comprensión de las ideas del autor , al hacer estás más claras y precisas.
¿Cómo ubicar la idea principal en el mismo?
Entre las ideas expuestas en un párrafo suele haber una más importante a la que
el autor se dedica a aclarar. Esta es la idea principal. Es muy importante
ubicarla porque ella nos permite percibir las relaciones entre las demás ideas u
nos ayuda a comprender el desarrollo del pensamiento del autor. Las otras ideas
que así contribuyen a este proceso son las llamadas ideas complementarias , pues
ellas se subordinan a la idea principal y la complementan y apoyan.
Donde ubicar esa idea principal?
Es muy frecuente hallarla al principio, en medio o al final de un párrafo, pero
esto no es una regla fija, puesto que en algunos casos la podemos encontrar
dividida ; una parte de la primera oración se encuentra en la primera oración y
la otra que comprende un segundo aspecto de ella , al final. Puede estar
distribuida en todas las oraciones del párrafo. Pero este caso es muy raro y
difícil. Utilizando las estrategias descritas a continuación se logrará identificarlas
sin mayores dificultades:
1. la oración mas general o más abstracta del párrafo.
2. La repetición de una idea en el texto.
3. La repetición de un concepto, mediante la misma palabra , un sinónimo o un
vocablo substituto.
4. Las palabras señal que señalan ideas generales, específicamente las que
introducen ideas terminales del razonamiento del autor.
5. El título del párrafo.
Todos estos indicios o algunos de ellos indican a menudo las ideas más
importantes y les pueden ayudar a deducir la idea principal del párrafo.
IDENTIFICACIÓN DE LAS IDEAS SECUNDARIAS O
COMPLEMENTARIAS
Como ubicar las ideas secundarias o complementarias?
Continuando con el estudio de las diferentes estrategias para identificar la idea
principal, presentaremos ahora las ideas complementarias, las cuales al ser
descartadas nos llevan a la idea principal sin que esta sufra distorsión alguna.
Las ideas complementarias se subordinan a la idea principal y la apoyan y
complementan.
Generalmente se encuentran:
1. en las oraciones de ejemplos y de comparación y contraste las cuales sirven
para ilustrar algún aspecto de la idea principal y el suprimirlas, esta no pierde su
sentido general.
2. Vienen introducidas por las palabras señal de pausa o retroceso.
3. suelen estar entre comas , guiones y paréntesis.
Veamos el siguiente ejercicio de ejemplo:
___________________________________

1).Integrated circuits (ICs) are probably the most important .components in electronics
today. An IC is a complete electronic circuit, often no larger than a discrete transistor,
integrated within a single chip of silicon. This type of IC is known as monolithic IC. It can
contain hundreds of transistors, diodes, resistors and capacitors, and all its components
an their connections are manufactured in the same production process. Monolithic
integrated circuits have substantial advantages over discrete circuits. One major
advantage of (ICs) is their small size.
2)Thanks to miniature circuitry, complex systems such as computers have reduced their
physical dimensions Low cost is another advantage of integrated circuits. Many circuits
are manufactured simultaneously. Consequently the cost of each circuit is lower. The
cost of processing a single semiconductor chip is then comparable to that of single
transistor. But perhaps the main advantage of integrated circuits is their high reliability i.e.
they fail less often than a discrete version of the same system because they require far
fewer solder joints and mechanical connections. Therefore ICs usually operate for longer
periods of time without a breakdown.
Topic:________________________________________
Idea principal:_________________________________________________________
Ideas secundarias:_____________________________________________________
Texto 2
3)A distinction is currently made between five degrees of integration or density - i.e. the
number of the elements or gate functions per mm2 of silicon area Small – Scale
Integration (SSI), Medium-Scale integration (MSI), Large- Scale Integration (LSI)Very
Large-Scale integration (VLSI), Ultra -Large-Scale integration (ULSI).
4) In SSi, a single integrated circuit chip includes an average number of a hundred
transistors and the average area needed is 3mm2. It has a number of logic functions that
ranges from 2 to 20.MSI circuits contain an average number of 500 transistors and the
average area needed is 8 mm2. The number of functions ranges from 20 to 100.
5)In LSI ,the average number of transistors is 100.000 and the average area 20 mm2
The number of functions ranges from 50.000 to 100.000. VLSI includes 250.000
transistors within an average area of 40 mm2. Functions range from 50.000
to100.000.The number of transistors in ULSI is 1.000.000 and the area, 70 mm2. The
number of functions can vary from 100.000 to 1.000.000.
Topic:________________________________________
Idea principal:_________________________________________________________
Ideas secundarias:_____________________________________________________
While the binary numeration system is an interesting mathematical abstraction, we
haven't yet seen its practical application to electronics. This chapter is devoted to just
that: practically applying the concept of binary bits to circuits. What makes binary
numeration so important to the application of digital electronics is the ease in which bits
may be represented in physical terms. Because a binary bit can only have one of two
different values, either 0 or 1, any physical medium capable of switching between two
saturated states may be used to represent a bit. Consequently, any physical system
capable of representing binary bits is able to represent numerical quantities, and
potentially has the ability to manipulate those numbers. This is the basic concept
underlying digital computing.

Electronic circuits are physical systems that lend themselves well to the representation
of binary numbers. Transistors, when operated at their bias limits, may be in one of two
different states: either cutoff (no controlled current) or saturation (maximum controlled
current). If a transistor circuit is designed to maximize the probability of falling into
either one of these states (and not operating in the linear, or active, mode), it can serve
as a physical representation of a binary bit. A voltage signal measured at the output of
such a circuit may also serve as a representation of a single bit, a low voltage
representing a binary "0" and a (relatively) high voltage representing a binary "1." Note
the following transistor circuit:
In this circuit, the transistor is in a state of saturation by virtue of the applied input
voltage (5 volts) through the two-position switch. Because it's saturated, the transistor drops
very little voltage between collector and emitter, resulting in an output voltage of
(practically) 0 volts. If we were using this circuit to represent binary bits, we would say that
the input signal is a binary "1" and that the output signal is a binary "0." Any voltage close
to full supply voltage (measured in reference to ground, of course) is considered a "1" and a
lack of voltage is considered a "0." Alternative terms for these voltage levels are high (same
as a binary "1") and low (same as a binary "0"). A general term for the representation of a
binary bit by a circuit voltage is logic level.

Moving the switch to the other position, we apply a binary "0" to the input and receive a
binary "1" at the output:
What we've created here with a single transistor is a circuit generally known as a logic
gate, or simply gate. A gate is a special type of amplifier circuit designed to accept and
generate voltage signals corresponding to binary 1's and 0's. As such, gates are not intended
to be used for amplifying analog signals (voltage signals between 0 and full voltage). Used
together, multiple gates may be applied to the task of binary number storage (memory
circuits) or manipulation (computing circuits), each gate's output representing one bit of a
multi-bit binary number. Just how this is done is a subject for a later chapter. Right now it is
important to focus on the operation of individual gates.

The gate shown here with the single transistor is known as an inverter, or NOT gate,
because it outputs the exact opposite digital signal as what is input. For convenience, gate
circuits are generally represented by their own symbols rather than by their constituent
transistors and resistors. The following is the symbol for an inverter:

An alternative symbol for an inverter is shown here:


Notice the triangular shape of the gate symbol, much like that of an operational
amplifier. As was stated before, gate circuits actually are amplifiers. The small circle, or
"bubble" shown on either the input or output terminal is standard for representing the
inversion function. As you might suspect, if we were to remove the bubble from the gate
symbol, leaving only a triangle, the resulting symbol would no longer indicate inversion,
but merely direct amplification. Such a symbol and such a gate actually do exist, and it is
called a buffer, the subject of the next section.

Like an operational amplifier symbol, input and output connections are shown as single
wires, the implied reference point for each voltage signal being "ground." In digital gate
circuits, ground is almost always the negative connection of a single voltage source (power
supply). Dual, or "split," power supplies are seldom used in gate circuitry. Because gate
circuits are amplifiers, they require a source of power to operate. Like operational
amplifiers, the power supply connections for digital gates are often omitted from the
symbol for simplicity's sake. If we were to show all the necessary connections needed for
operating this gate, the schematic would look something like this:

Power supply conductors are rarely shown in gate circuit schematics, even if the power
supply connections at each gate are. Minimizing lines in our schematic, we get this:

"Vcc" stands for the constant voltage supplied to the collector of a bipolar junction
transistor circuit, in reference to ground. Those points in a gate circuit marked by the label
"Vcc" are all connected to the same point, and that point is the positive terminal of a DC
voltage source, usually 5 volts.

As we will see in other sections of this chapter, there are quite a few different types of
logic gates, most of which have multiple input terminals for accepting more than one
signal. The output of any gate is dependent on the state of its input(s) and its logical
function.

One common way to express the particular function of a gate circuit is called a truth
table. Truth tables show all combinations of input conditions in terms of logic level states
(either "high" or "low," "1" or "0," for each input terminal of the gate), along with the
corresponding output logic level, either "high" or "low." For the inverter, or NOT, circuit
just illustrated, the truth table is very simple indeed:

Truth tables for more complex gates are, of course, larger than the one shown for the
NOT gate. A gate's truth table must have as many rows as there are possibilities for unique
input combinations. For a single-input gate like the NOT gate, there are only two
possibilities, 0 and 1. For a two input gate, there are four possibilities (00, 01, 10, and 11),
and thus four rows to the corresponding truth table. For a three-input gate, there are eight
possibilities (000, 001, 010, 011, 100, 101, 110, and 111), and thus a truth table with eight
rows are needed. The mathematically inclined will realize that the number of truth table
rows needed for a gate is equal to 2 raised to the power of the number of input terminals.

 REVIEW:
 In digital circuits, binary bit values of 0 and 1 are represented by voltage
signals measured in reference to a common circuit point called ground. An absence
of voltage represents a binary "0" and the presence of full DC supply voltage
represents a binary "1."
 A logic gate, or simply gate, is a special form of amplifier circuit designed to
input and output logic level voltages (voltages intended to represent binary bits).
Gate circuits are most commonly represented in a schematic by their own unique
symbols rather than by their constituent transistors and resistors.
 Just as with operational amplifiers, the power supply connections to gates
are often omitted in schematic diagrams for the sake of simplicity.
A truth table is a standard way of representing the input/output relationships of a gate
circuit, listing all the possible input logic level combinations with their respective output
logic levels.
INSTITUTO UNIVERSITARIO DE TECNOLOGÍA
AGROINDUSTRIAL

DIGITAL SIGNALS AND GATES

IDEAS PRINCIPALES Y SECUNDARIAS

Prof. Elsa Rangel de Zambrano


SAN CRISTÒBAL, ______, _______, ______

ACTIVITY 2 (score: 5 points)

1) Halle el tópico o tema central del texto. ____________________________________

2) Determine los sub- temas o tópicos que el autor desarrolla en el tema.

1_______________________________________________________________________________________

2_______________________________________________________________________________________

3______________________________________________________________________________________

4______________________________________________________________________________________

5______________________________________________________________________________________

6_____________________________________________________________________________________

3) ¿Que ideas apoyan las ideas centrales desarrolladas por el autor en cada uno de los
párrafos?

FULLNAME:
_____________________________
_____________________________
Score:_____________________

Probablemente la habilidad de comprensión mas importante que los lectores


pueden desarrollar es el reconocimiento de los patrones organizacionales de
un texto. Las investigaciones han demostrado (Carell 1984, 1985; Meyer and
Rice 1987) que los lectores comprenden y recuerdan más los materiales que
están organizacionalmente claros para ellos y que diferentes modos
organizacionales varían en la accesibilidad para los hablantes de diferentes
idiomas (Carell 1985).
Existen diferentes tipos de oraciones, entre las mas comunes están las ideas
principales y las ideas secundarias, también observamos en el objetivo anterior que
hay palabras señal que indican en un momento dado lo que el autor desea
expresar bien sea una comparación o contraste, una idea de proceso o
información específica sobre determinado tema. Basados en estos elementos , que
son las categorías de oraciones y las palabras señal daremos comienzo al análisis de
la estructura del párrafo.

El párrafo esta compuesto de diferentes categorías de oraciones estas están


clasificadas de acuerdo a su contenido.
1. Información específica
Las palabras señal muestran alguna información específica, tal como
ejemplos y referencias particulares a un tema específico.
LED2 constitutes an inexpensive but effective voltaje reference , which imposes a
constant voltaje across current –setting resistorR 1 provided that you choose R3’s
value to suit Q2’s base drive , you can set the current in LED 2 and the voltage across
it to fairly precise and constant values. For example, with R1= 430 , the current in
LED2 is approximately 10 mA with 5V at Q2 ‘s base (Tx high).
2. Ideas generales y detalles
Con frecuencia un autor lista unos pocos hechos acerca de un tema quizás
para explicar una idea principal , sustentar una opinión o dar una evidencia
acerca de una oración en general. En Inglés esta es una estructura de
organización muy común. Es con frecuencia llamado estructura de listado.

The result of an imbalance of this "fluid" (electrons) between objects is


called static electricity. It is called "static" because the displaced electrons
tend to remain stationary after being moved from one material to another.
In the case of wax and wool, it was determined through further
experimentation that electrons in the wool actually transferred to the atoms
in the wax, which is exactly opposite of Franklin's conjecture! In honour of
Franklin's designation of the wax's charge being "negative" and the wool's
charge being "positive," electrons are said to have a "negative" charging
influence. Thus, an object whose atoms have received a surplus of
electrons is said to be negatively charged, while an object whose atoms are
lacking electrons is said to be positively charged, as confusing as these
designations may seem. By the time the true nature of electric "fluid" was
discovered, Franklin's nomenclature of electric charge was too well
established to be easily changed, and so it remains to this day.

3. De organización espacial
La organización espacial es usada para describir la localización de lugares,
objetos y gente en el espacio. Es también usada para describir movimiento a
través del espacio. Es usada por los ingenieros para describir las partes de
un motor, los electricistas para determinar la organización de un circuito, los
historiadores para describir las rutas seguidas por los exploradores y los
arquitectos para describir las edificaciones. Un párrafo con organización
espacial permite a los lectores formarse una imagen de lo que ellos leen.

The tube is full of marbles, just as a conductor is full of free electrons


ready to be moved by an outside influence. If a single marble is suddenly
inserted into this full tube on the left-hand side, another marble will
immediately try to exit the tube on the right. Even though each marble
only travelled a short distance, the transfer of motion through the tube is
virtually instantaneous from the left end to the right end, no matter how
long the tube is. With electricity, the overall effect from one end of a
conductor to the other happens at the speed of light: a swift 186,000 miles
per second!!! Each individual electron though, travels through the
conductor at a much slower pace.

4. De contraste:
Un párrafo de contraste muestra diferencias . Un escritor puede contrastar
personas, lugares, objetos u eventos. Por ejemplo el contraste es usado por
lo lingüistas para presentar diferencias entre los idiomas , por los
periodistas entre candidatos a cargos públicos, por antropólogos para
presentar diferencias entre diferentes culturas, los electricistas para mostrar
diferencias entre un dispositivo y otro, entre aparatos electrónicos , etc.

Neutrons are much less influential on the chemical character and identity of an atom
than protons, although they are just as hard to add to or remove from the nucleus,
being so tightly bound. If neutrons are added or gained, the atom will still retain the
same chemical identity, but its mass will change slightly and it may acquire strange
nuclear properties such as radioactivity.

However, electrons have significantly more freedom to move around in an


atom than either protons or neutrons. In fact, they can be knocked out of
their respective positions (even leaving the atom entirely!) by far less energy
than what it takes to dislodge particles in the nucleus. If this happens, the
atom still retains its chemical identity, but an important imbalance occurs.
Electrons and protons are unique in the fact that they are attracted to one
another over a distance. It is this attraction over distance which causes
5. De orden cronológico:
Un capítulo que describe la historia de una máquina , una idea, una
persona o un suceso es generalmente escrito en orden cronológico. Por
ejemplo un capítulo acerca del desarrollo del radio, de la vida de J. F.
Kennedy o la historia de la Segunda Guerra Mundial estaría escrito en
orden cronológico.

Electronics began at the start of the twentieth century with the invention
of the vacuum tube. The first devices for everyday use were radios,
followed by televisions , records players , and tape recorders. These devices
were large and used a lot of power.
The invention of the transistor in 1947 meant that much smaller ,low-
powered devices could be developed. A wide variety of electronic devices
such as hi-fi units and portable radios became common in the home.
It was not until 1958 that microelectronics s began with the development
of ICs ( integrated Circuits) on silicon chips. This led to a great increase
in the use of electronics in everyday items.
Nowdays, microprocessor are used to control many household items such
as automatic washing –machines, dishwashers, central heating systems,
sewing machines and food processors.
In the future, electronics are likely to become even more common in the
home as multimedia entertainment systems and computer-controlled
robots are developed.

6. De clasificación
You can think of a transistor as two back-to-back diodes in one package as
Unshown in Figure
párrafo 5. Click Here
de clasificación For ASCII
divide Version
personas, of Figure
lugares 5. en grupos . La
o cosas
clasificación es usada para describir diferentes categorías.

Note that transistors come in two basic types: NPN and PNP. The letters C,
B, E stand for COLLECTOR, BASE, EMITTER which are the names of the
three leads which come out of a transistor.

Transistors come in many different case styles, three of which are shown in
Figure 6. It is important to know where C, B, E are for any given case. Click
Here For ASCII Version of Figure 6.
7. Proceso
Un párrafo de proceso describe como hacer algo o como una máquina o
sistema funciona. Por ejemplo párrafos que describen como hacer un
experimento o como el motor de un automóvil funciona son párrafos acerca
de procesos.

First, ensure that the transmitter is as tightly constructed as possible, with


very short wires and no dodgy joints. A PCB could be a good idea (TE sells
one that would nearly suit, i.e. the Voyager) but I have had good results with
point-to-point wiring on a wire-wrap board (NOT using wire-wrap as this
introduces excess stray inductance) Make sure that there is adequate
capacitance across the supply rails, particularly if the transmitter shares a
power supply with a receiver. Second, Consider some kind of shielding - a
sheet of aluminum foil with contact on both sides can be used to make a
decent shield. Connect the finished article to a spare serial port. Assuming
you are running Linux (of course you are), do something like "yes 'Hello
World' > /dev/cua1", after that, setting the serial port to the desired baud
rate. For anyone unfortunate enough to be using a Microsoft OS, perhaps
you could use Telex or some such commands program to send stuff to the
port.

8. De causa y efecto
Los autores usan oraciones de causa y efecto en las siguientes formas :
a. Para decir que una acción causa otra.
Ejemplo: (cause) The bookstore announced a special sale on textbook yesterday
(Effect) Therefore we rushed to the store to buy the books we needed

b. Para presentar razones para una acción.


Ejemplo: (reason) I want to save money for my vacation
(conclusion) so, I will not spend any money on the movies this week.

c. Para dar propósitos u objetivos:

Ejemplo:

I saved my money For a vacation.


To go on a vacation.
In order to go on a vacation.
So that I could go on a vacation.

d. Para presentar condiciones y resultados:

Ejemplo: (condition) If I spend all my money now,


(result) I will not be able to travel next summer.

(condition) If I saved money this year,


(result) I would travel next year.

Presentar rezones para una acción


Following Franklin's speculation of the wool rubbing something off of
the wax, the type of charge that was associated with rubbed wax became
known as "negative" (because it was supposed to have a deficiency of
fluid) while the type of charge associated with the rubbing wool became
known as "positive" (because it was supposed to have an excess of
fluid). Little did he know that his innocent conjecture would cause much
confusion for students of electricity in the future!

Condición y resultado
Electrons tend to repel other electrons over a distance, as do protons
with other protons. The only reason protons bind together in the
nucleus of an atom is because of a much stronger force called the
strong nuclear force which has effect only under very short distances.
Because of this attraction/repulsion behaviour between individual
particles, electrons and protons are said to have opposite electric
charges. That is, each electron has a negative charge, and each proton a
positive charge. In equal numbers within an atom, they counteract each
other's presence so that the net charge within the atom is zero. This is
why the picture of a carbon atom had six electrons: to balance out the
electric charge of the six protons in the nucleus.
1-Magnetic fields and inductance

a-Whenever electrons flow through a conductor, a magnetic field will develop


around that conductor. This effect is called electromagnetism. Magnetic fields
effect the alignment of electrons in an atom, and can cause physical force to
develop between atoms across space just as with electric fields developing force
between electrically charged particles. Like electric fields, magnetic fields can
occupy completely empty space, and affect matter at a distance.

b-Fields have two measures: a field force and a field flux. The field force is the
amount of "push" that a field exerts over a certain distance. The field flux is the
total quantity, or effect, of the field through space. Field force and flux are roughly
analogous to voltage ("push") and current (flow) through a conductor,
respectively, although field flux can exist in totally empty space (without the
motion of particles such as electrons) whereas current can only take place where
there are free electrons to move. Field flux can be opposed in space, just as the
flow of electrons can be opposed by resistance. The amount of field flux that will
develop in space is proportional to the amount of field force applied, divided by
the amount of opposition to flux. Just as the type of conducting material dictates
that conductor's specific resistance to electric current, the type of material
occupying the space through which a magnetic field force is impressed dictates
the specific opposition to magnetic field flux.

c-Whereas an electric field flux between two conductors allows for an


accumulation of free electron charge within those conductors, an electromagnetic
field flux allows for a certain "inertia" to accumulate in the flow of electrons
through the conductor producing the field.

d-Inductors are components designed to take advantage of this phenomenon by


shaping the length of conductive wire in the form of a coil. This shape creates a
stronger magnetic field than what would be produced by a straight wire. Some
inductors are formed with wire wound in a self-supporting coil. Others wrap the
wire around a solid core material of some type. Sometimes the core of an inductor
will be straight, and other times it will be joined in a loop (square, rectangular, or
circular) to fully contain the magnetic flux. These design options all have effect on
the performance and characteristics of inductors.

e-The schematic symbol for an inductor, like the capacitor, is quite simple, being little
more than a coil symbol representing the coiled wire. Although a simple coil shape is
the generic symbol for any inductor, inductors with cores are sometimes distinguished
by the addition of parallel lines to the axis of the coil. A newer version of the inductor
symbol dispenses with the coil shape in favor of several "humps" in a row:

f-As the electric current produces a concentrated magnetic field around the coil, this
field flux equates to a storage of energy representing the kinetic motion of the
electrons through the coil. The more current in the coil, the stronger the magnetic field
will be, and the more energy the inductor will store.

g-Because inductors store the kinetic energy of moving electrons in the form of a
magnetic field, they behave quite differently than resistors (which simply dissipate
energy in the form of heat) in a circuit. Energy storage in an inductor is a function of
the amount of current through it. An inductor's ability to store energy as a function of
current results in a tendency to try to maintain current at a constant level. In other
words, inductors tend to resist changes in current. When current through an inductor
is increased or decreased, the inductor "resists" the change by producing a voltage
between its leads in opposing polarity to the change.
h-To store more energy in an inductor, the current through it must be increased. This
means that its magnetic field must increase in strength, and that change in field
strength produces the corresponding voltage according to the principle of
electromagnetic self-induction. Conversely, to release energy from an inductor, the
current through it must be decreased. This means that the inductor's magnetic field
must decrease in strength, and that change in field strength self-induces a voltage drop
of just the opposite polarity.

Just as Isaac Newton's first Law of Motion ("an object in motion tends to stay in motion; an
object at rest tends to stay at rest") describes the tendency of a mass to oppose changes in
velocity, we can state an inductor's tendency to oppose changes in current as such:
"Electrons moving through an inductor tend to stay in motion; electrons at rest in an
inductor tend to stay at rest." Hypothetically, an inductor left short-circuited will maintain a
constant rate of current through it with no external assistance:

Practically speaking, however, the ability for an inductor to self-sustain current is realized
only with superconductive wire, as the wire resistance in any normal inductor is enough to
cause current to decay very quickly with no external source of power.

When the current through an inductor is increased, it drops a voltage opposing the direction
of electron flow, acting as a power load. In this condition the inductor is said to be
charging, because there is an increasing amount of energy being stored in its magnetic
field. Note the polarity of the voltage with regard to the direction of current:

Conversely, when the current through the inductor is decreased, it drops a voltage aiding
the direction of electron flow, acting as a power source. In this condition the inductor is said
to be discharging, because its store of energy is decreasing as it releases energy from its
magnetic field to the rest of the circuit. Note the polarity of the voltage with regard to the
direction of current.
If a source of electric power is suddenly applied to an unmagnetized inductor, the inductor
will initially resist the flow of electrons by dropping the full voltage of the source. As
current begins to increase, a stronger and stronger magnetic field will be created, absorbing
energy from the source. Eventually the current reaches a maximum level, and stops
increasing. At this point, the inductor stops absorbing energy from the source, and is
dropping minimum voltage across its leads, while the current remains at a maximum level.
As an inductor stores more energy, its current level increases, while its voltage drop
decreases. Note that this is precisely the opposite of capacitor behavior, where the storage
of energy results in an increased voltage across the component! Whereas capacitors store
their energy charge by maintaining a static voltage, inductors maintain their energy
"charge" by maintaining a steady current through the coil.

The type of material the wire is coiled around greatly impacts the strength of the magnetic
field flux (and therefore how much stored energy) generated for any given amount of
current through the coil. Coil cores made of ferromagnetic materials (such as soft iron) will
encourage stronger field fluxes to develop with a given field force than nonmagnetic
substances such as aluminum or air.

The measure of an inductor's ability to store energy for a given amount of current flow is
called inductance. Not surprisingly, inductance is also a measure of the intensity of
opposition to changes in current (exactly how much self-induced voltage will be produced
for a given rate of change of current). Inductance is symbolically denoted with a capital
"L," and is measured in the unit of the Henry, abbreviated as "H."

An obsolete name for an inductor is choke, so called for its common usage to block
("choke") high-frequency AC signals in radio circuits. Another name for an inductor, still
used in modern times, is reactor, especially when used in large power applications. Both of
these names will make more sense after you've studied alternating current (AC) circuit
theory, and especially a principle known as inductive reactance.

2-Inductors and calculus

Inductors do not have a stable "resistance" as conductors do. However, there is a definite
mathematical relationship between voltage and current for an inductor, as follows:
You should recognize the form of this equation from the capacitor chapter. It relates one
variable (in this case, inductor voltage drop) to a rate of change of another variable (in this
case, inductor current). Both voltage (v) and rate of current change (di/dt) are
instantaneous: that is, in relation to a specific point in time, thus the lower-case letters "v"
and "i". As with the capacitor formula, it is convention to express instantaneous voltage as
v rather than e, but using the latter designation would not be wrong. Current rate-of-change
(di/dt) is expressed in units of amps per second, a positive number representing an increase
and a negative number representing a decrease.

Like a capacitor, an inductor's behavior is rooted in the variable of time. Aside from any
resistance intrinsic to an inductor's wire coil (which we will assume is zero for the sake of
this section), the voltage dropped across the terminals of an inductor is purely related to
how quickly its current changes over time.

Suppose we were to connect a perfect inductor (one having zero ohms of wire resistance) to
a circuit where we could vary the amount of current through it with a potentiometer
connected as a variable resistor:

If the potentiometer mechanism remains in a single position (wiper is stationary), the


series-connected ammeter will register a constant (unchanging) current, and the voltmeter
connected across the inductor will register 0 volts. In this scenario, the instantaneous rate of
current change (di/dt) is equal to zero, because the current is stable. The equation tells us
that with 0 amps per second change for a di/dt, there must be zero instantaneous voltage (v)
across the inductor. From a physical perspective, with no current change, there will be a
steady magnetic field generated by the inductor. With no change in magnetic flux (dΦ/dt =
0 Webers per second), there will be no voltage dropped across the length of the coil due to
induction.
If we move the potentiometer wiper slowly in the "up" direction, its resistance from end to
end will slowly decrease. This has the effect of increasing current in the circuit, so the
ammeter indication should be increasing at a slow rate:

Assuming that the potentiometer wiper is being moved such that the rate of current increase
through the inductor is steady, the di/dt term of the formula will be a fixed value. This fixed
value, multiplied by the inductor's inductance in Henrys (also fixed), results in a fixed
voltage of some magnitude. From a physical perspective, the gradual increase in current
results in a magnetic field that is likewise increasing. This gradual increase in magnetic flux
causes a voltage to be induced in the coil as expressed by Michael Faraday's induction
equation e = N(dΦ/dt). This self-induced voltage across the coil, as a result of a gradual
change in current magnitude through the coil, happens to be of a polarity that attempts to
oppose the change in current. In other words, the induced voltage polarity resulting from an
increase in current will be oriented in such a way as to push against the direction of
current, to try to keep the current at its former magnitude. This phenomenon exhibits a
more general principle of physics known as Lenz's Law, which states that an induced effect
will always be opposed to the cause producing it.

In this scenario, the inductor will be acting as a load, with the negative side of the induced
voltage on the end where electrons are entering, and the positive side of the induced voltage
on the end where electrons are exiting.
Changing the rate of current increase through the inductor by moving the potentiometer
wiper "up" at different speeds results in different amounts of voltage being dropped across
the inductor, all with the same polarity (opposing the increase in current):

Here again we see the derivative function of calculus exhibited in the behavior of an
inductor. In calculus terms, we would say that the induced voltage across the inductor is the
derivative of the current through the inductor: that is, proportional to the current's rate-of-
change with respect to time.

Reversing the direction of wiper motion on the potentiometer (going "down" rather than
"up") will result in its end-to-end resistance increasing. This will result in circuit current
decreasing (a negative figure for di/dt). The inductor, always opposing any change in
current, will produce a voltage drop opposed to the direction of change:
How much voltage the inductor will produce depends, of course, on how rapidly the current
through it is decreased. As described by Lenz's Law, the induced voltage will be opposed to
the change in current. With a decreasing current, the voltage polarity will be oriented so as
to try to keep the current at its former magnitude. In this scenario, the inductor will be
acting as a source, with the negative side of the induced voltage on the end where electrons
are exiting, and the positive side of the induced voltage on the end where electrons are
entering. The more rapidly current is decreased, the more voltage will be produced by the
inductor, in its release of stored energy to try to keep current constant.

Again, the amount of voltage across a perfect inductor is directly proportional to the rate of
current change through it. The only difference between the effects of a decreasing current
and an increasing current is the polarity of the induced voltage. For the same rate of current
change over time, either increasing or decreasing, the voltage magnitude (volts) will be the
same. For example, a di/dt of -2 amps per second will produce the same amount of induced
voltage drop across an inductor as a di/dt of +2 amps per second, just in the opposite
polarity.

If current through an inductor is forced to change very rapidly, very high voltages will be
produced. Consider the following circuit:

In this circuit, a lamp is connected across the terminals of an inductor. A switch is used to
control current in the circuit, and power is supplied by a 6 volt battery. When the switch is
closed, the inductor will briefly oppose the change in current from zero to some magnitude,
but will drop only a small amount of voltage. It takes about 70 volts to ionize the neon gas
inside a neon bulb like this, so the bulb cannot be lit on the 6 volts produced by the battery,
or the low voltage momentarily dropped by the inductor when the switch is closed:

When the switch is opened, however, it suddenly introduces an extremely high resistance
into the circuit (the resistance of the air gap between the contacts). This sudden introduction
of high resistance into the circuit causes the circuit current to decrease almost instantly.
Mathematically, the di/dt term will be a very large negative number. Such a rapid change of
current (from some magnitude to zero in very little time) will induce a very high voltage
across the inductor, oriented with negative on the left and positive on the right, in an effort
to oppose this decrease in current. The voltage produced is usually more than enough to
light the neon lamp, if only for a brief moment until the current decays to zero:

For maximum effect, the inductor should be sized as large as possible (at least 1 Henry of
inductance).

3-Factors affecting inductance

There are four basic factors of inductor construction determining the amount of inductance
created. These factors all dictate inductance by affecting how much magnetic field flux will
develop for a given amount of magnetic field force (current through the inductor's wire
coil):
NUMBER OF WIRE WRAPS, OR "TURNS" IN THE COIL: All other factors being
equal, a greater number of turns of wire in the coil results in greater inductance; fewer turns
of wire in the coil results in less inductance.

Explanation: More turns of wire means that the coil will generate a greater amount of
magnetic field force (measured in amp-turns!), for a given amount of coil current.

COIL AREA: All other factors being equal, greater coil area (as measured looking
lengthwise through the coil, at the cross-section of the core) results in greater inductance;
less coil area results in less inductance.

Explanation: Greater coil area presents less opposition to the formation of magnetic field
flux, for a given amount of field force (amp-turns).

COIL LENGTH: All other factors being equal, the longer the coil's length, the less
inductance; the shorter the coil's length, the greater the inductance.

Explanation: A longer path for the magnetic field flux to take results in more opposition to
the formation of that flux for any given amount of field force (amp-turns).
CORE MATERIAL: All other factors being equal, the greater the magnetic permeability
of the core which the coil is wrapped around, the greater the inductance; the less the
permeability of the core, the less the inductance.

Explanation: A core material with greater magnetic permeability results in greater magnetic
field flux for any given amount of field force (amp-turns).

An approximation of inductance for any coil of wire can be found with this formula:

It must be understood that this formula yields approximate figures only. One reason for this
is the fact that permeability changes as the field intensity varies (remember the nonlinear
"B/H" curves for different materials). Obviously, if permeability (µ) in the equation is
unstable, then the inductance (L) will also be unstable to some degree as the current
through the coil changes in magnitude. If the hysteresis of the core material is significant,
this will also have strange effects on the inductance of the coil. Inductor designers try to
minimize these effects by designing the core in such a way that its flux density never
approaches saturation levels, and so the inductor operates in a more linear portion of the
B/H curve.

If an inductor is designed so that any one of these factors may be varied at will, its
inductance will correspondingly vary. Variable inductors are usually made by providing a
way to vary the number of wire turns in use at any given time, or by varying the core
material (a sliding core that can be moved in and out of the coil). An example of the former
design is shown in this photograph:
This unit uses sliding copper contacts to tap into the coil at different points along its length.
The unit shown happens to be an air-core inductor used in early radio work.

A fixed-value inductor is shown in the next photograph, another antique air-core unit built
for radios. The connection terminals can be seen at the bottom, as well as the few turns of
relatively thick wire:

Here is another inductor (of greater inductance value), also intended for radio applications.
Its wire coil is wound around a white ceramic tube for greater rigidity:

Inductors can also be made very small for printed circuit board applications. Closely
examine the following photograph and see if you can identify two inductors near each
other:
The two inductors on this circuit board are labeled L 1 and L2, and they are located to the
right-center of the board. Two nearby components are R3 (a resistor) and C16 (a capacitor).
These inductors are called "toroidal" because their wire coils are wound around donut-
shaped ("torus") cores.

Like resistors and capacitors, inductors can be packaged as "surface mount devices" as
well. The following photograph shows just how small an inductor can be when packaged as
such:

A pair of inductors can be seen on this circuit board, to the right and center, appearing as
small black chips with the number "100" printed on both. The upper inductor's label can be
seen printed on the green circuit board as L5. Of course these inductors are very small in
inductance value, but it demonstrates just how tiny they can be manufactured to meet
certain circuit design needs.

4-Series and parallel inductors

When inductors are connected in series, the total inductance is the sum of the individual
inductors' inductances. To understand why this is so, consider the following: the definitive
measure of inductance is the amount of voltage dropped across an inductor for a given rate
of current change through it. If inductors are connected together in series (thus sharing the
same current, and seeing the same rate of change in current), then the total voltage dropped
as the result of a change in current will be additive with each inductor, creating a greater
total voltage than either of the individual inductors alone. Greater voltage for the same rate
of change in current means greater inductance.
Thus, the total inductance for series inductors is more than any one of the individual
inductors' inductances. The formula for calculating the series total inductance is the same
form as for calculating series resistances:

When inductors are connected in parallel, the total inductance is less than any one of the
parallel inductors' inductances. Again, remember that the definitive measure of inductance
is the amount of voltage dropped across an inductor for a given rate of current change
through it. Since the current through each parallel inductor will be a fraction of the total
current, and the voltage across each parallel inductor will be equal, a change in total current
will result in less voltage dropped across the parallel array than for any one of the inductors
considered separately. In other words, there will be less voltage dropped across parallel
inductors for a given rate of change in current than for any of of those inductors considered
separately, because total current divides among parallel branches. Less voltage for the same
rate of change in current means less inductance.

Thus, the total inductance is less than any one of the individual inductors' inductances. The
formula for calculating the parallel total inductance is the same form as for calculating
parallel resistances:

 REVIEW:
 Inductances add in series.
 Inductances diminish in parallel.
INSTITUTO UNIVERSITARIO DE TECNOLOGÍA
AGROINDUSTRIAL

INDUCTORS

ESTRUCTURA DEL PARRAFO

Prof. Elsa Rangel de Zambrano

SAN CRISTÒBAL, ______, _______, ______


Activity 3 (5 points)

1-Determina el tema o tópico de la lectura:____________________________________


2-Extraiga las ideas centrales desarrolladas en toda la lectura.
a)______________________________________________________________________
b)______________________________________________________________________
c)______________________________________________________________________
d)______________________________________________________________________
3-Con el texto anterior realize las siguientes actividades:

Determina la estructura predominante en el párrafo 1 letras e,f,g,h.


Párrafo e: _______________________________________________________________
Párrafo f: _______________________________________________________________
Párrafo g: _______________________________________________________________
Párrafo h: _______________________________________________________________

4- responda las siguientes preguntas:

1- What are inductors?_____________________________________________________


________________________________________________________________________
________________________________________________________________________
2- How should the current be to store more energy in a inductor?__________________
_________________________________________________________________________
_________________________________________________________________________

3-How is the measure of an inductor’s ability to store energy called?__________________


_________________________________________________________________________
_________________________________________________________________________

4-What happens if the potentiometer mechanism remains in a single position?_______


_________________________________________________________________________

5-What are the four basic factors of construction determining the amount of
inductance created?
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
6- What happens when inductors are connected in parallel?________________________
_________________________________________________________________________

5-Traduzca los párrafos e,f,g,h


FULLNAME:
____________________ Score:_____________________
__________________

¿Qué es?Un esquema básico (outline) muestra una visión básica y detalles
importantes de cualquier tema descrito en un texto. Un esquema básico(outline)es:

 Una descripción lógica y general


 Un resumen esquemático
 Un patrón organizacional.
 Un diseño visual y conceptual del texto.
 Un esquema básico refleja un pensamiento lógico y una clara clasificación de
ideas principales y secundarias. En cualquiera de estos casos un esquema
básico(outline)mostrará una

¿Por que hacerlo?

Esto te ayudará a:

1. En el proceso de comprensión de un texto, organizando tus ideas al respecto.


2. Determinar las ideas principales y secundarias del mismo. Mostrando la relación
entre ideas.
3. Resolver el orden en el cual los tópicos u subtópicos aparecen en la lectura.
Presentándolo de una forma lógica.
4. Obtener una visión ordenada del texto.
5. Definir límites y grupos es decir temas y sub-temas

Un esquema básico puede ser usado para determinar que has entendido sobre la
lectura de un texto. Este esquema básico te da la oportunidad de centrarte en lo
más importante de un texto conduciéndote a una comprensión rápida y general del
mismo.

¿Como hacerlo?

1. Identifique el tópico: El tópico es importante. Trata de encontrar el punto principal


sobre lo que trata el texto. Resúmelo en una oración o frase corta.
2. Identifica las categorías principales . ¿Cuales son los puntos que se tratan en el
texto?
3. Determine la primera categoría . ¿Cual es el primer punto que se cubre?
4. Identifique las subcategorías: que ideas apoyan al punto principal . El número de
categorías depende de la cantidad de información que se maneje no hay un numero
correcto o incorrecto para usar las subcategorías. Cada categoría debe tener como
mínimo 2 entradas. Si la primera categoría es un número romano I tu esquema básico
debe tener una categoría con el numero romano II . Si tienes una letra mayúscula bajo
la categoría de ideas principales debes tener una categoría con la letra mayúscula B o
C, D, E dependiendo de la información que se cubra. Si hay un a sub-idea debajo de la
categoría de letras mayúsculas con números cardinales (1) tiene que haber una
subcategoría 2 y así sucesivamente, dependiendo de la cantidad de información que el
autor del texto haya cubierto en el. Aquí hay un ejemplo para ilustrar como se
construye un esquema básico.

ejemplo:

I. MAIN IDEA
A. Subsidiary idea or supporting idea to I
B. Subsidiary idea or supporting idea to I
1. Subsidiary idea to B
2. Subsidiary idea to B
a) Subsidiary idea to 2
b) Subsidiary idea to 2

II. MAIN IDEA


A. Subsidiary or supporting idea to II
B. Subsidiary idea to II
C. Subsidiary idea to II

III. MAIN IDEA

ESQUEMA BÁSICO CON UN TEMA

Supongamos que la lectura se trata de SIDA y determinamos algunas de las ideas


principales sobre las que el autor desarrolló el texto:

Major Aspects of Aids

I. Transmittal of AIDS
A. Transfusions
B. Body fluids
1. Sexual
2. Non-sexual
II. Societal Consequences of AIDS
A. Epidemic disease pattern
1. Teenagers
2. Women
3. Homosexuals
B. AIDS babies
C. Increased homophobia
D. Overburdened health care

III. Research Solutions to AIDS


A. AZT
B. HIV virus
C. Other viruses
1-Electricity is the flow of electrons in a conductor and there are four quite intuitive quantities help
to characterize it. voltage, current, resistance and power

The first is voltage. This term refers to the level of energy electrons have relative to some
reference point (often called ground in a circuit). The higher the voltage, the more energy electrons have
to do work as they travel through the circuit. In general, if two points are at a different voltage relative to
each other, electricity will flow from one to the other if they are connected by something that conducts
electricity. The unit of measurement of voltage is the Volt (V). As a sample of voltages our wall outlets at
home (at least in the US) are 110V, the AA, C & D cells we buy at the corner store are all rated at 1.5V,
and the electronics on Teleo Modules requires 5V.

The next quantity is current. This is an expression of how much charge is travelling through the conductor per
second. The unit of measurement for current is the Amp (A). You can see that voltage and current are separate
things: you can have a very small current at a very high voltage, a huge current at a very high voltage and so
on.

The next quantity is resistance. Resistance is an expression of the degree to which electron flow will be
impeded through a conductor. The unit is the Ohm ( ). In simple circuits resistance determines the relation
between voltage and current. At the extremes, a short piece of wire will have a resistance of nearly zero Ohms,
while an air gap (for example in an open switch) has very large resistance (millions of Ohms). Intuitively a
couple of relationships will hold: in a conductor, a voltage difference between the two ends will cause a current
to flow. How much current will be determined by how much resistance the conductor offers. If there's less
resistance more current will flow. In fact, given a power source of high enough capacity, if you half the
resistance, you will double the current. Conversely, if you double the resistance, you will half the current.

The final quantity is power. The unit of power is the Watt. It's an expression of the overall energy consumed by
a component. It is worked out by multiplying the voltage and the current together - P = VI. For example if a
motor was running at 12V and the current it was drawing was 2A, the power it would be dissipating would be
24W.
2- BASIC CIRCUITS

For electricity to flow, there needs to be a path that connects all the elements together. In the diagram below,
you can see how electricity can travel from the cell around in a loop through the lamp and back to the cell
again provided all the wires are in their proper places.

It should be noted that we show electricity travelling from the positive (+) side of the cell around the circuit to

the negative side. This is called conventional current. The slightly odd thing about this is that the
electrons that constitute electrical current are negatively charged and actually travel in the
opposite direction. The fact that we depict current travelling from the positive to the
negative is an historical accident. Fortunately unless you're doing something esoteric like
semiconductor physics, this extra layer of complexity need never worry you.

Lamp and Cell Circuit

Circuit diagrams provide a very efficient way to describe an electronic circuit. They use a small set of symbols
and conventions that need to be learned but the benefits of their form over a more pictorial style are so
definitive that they are used universally. The circuit above can be diagrammed more efficiently. In order to
illustrate this, here are some symbols used to depict cells and lamp.

Cell. This is often called a battery, but technically a battery is multiple


cells. This is what C, D, AA and AAA cells we can get at the corner store
all are. This kind of cells is rated at 1.5V.

Lamp. Like you might find in a flashlight. Lamps have voltage ratings like
many things. This rating indicates the voltage that the lamp is designed to
run at. It will be the highest voltage the lamp can withstand without
getting too hot and burning out. Lamps may also state their wattage - the
power they consume. From this and the a re-arrangement of the equation
for power (I = P / V) the likely current consumed can be calculated.

With these symbols we can now construct a circuit diagram with a cell and a lamp.
Simple Lamp and Cell Circuit Diagram

This circuit diagram expresses only the essential features of the circuit. The abstract symbols hide all the
myriad details of appearance that are actually irrelevant to the circuit. For example, the circuit is not influenced
by the colors of the wires so they're not indicated, curves in the wire are irrelevant so the lines in the diagram
are all straight, the actual physical design of the light bulb is from an electrical point of view, immaterial, etc.

The lines on the diagram need not even be wires. They can be traces of copper on a printed circuit board
(PCB) or pretty much anything that conducts. In a car, for example, wires that go from parts of the motor,
headlights and other devices back to the negative terminal of the battery are not needed. The metal of the car's
chassis and body are the return path to the battery and complete the circuit.

Another very important abstraction is the connection of wires. People are often confused by
this. The diagram below shows the same circuit as before except for the addition of another
lamp. The problems arise when people try to figure out exactly how to connect the things
together. How literally does the diagram need to be followed?

Dual Lamp Circuit

Take the top section of the circuit. It depicts a long section of conductor with one lamp
getting its power from it about half way along and another getting its power at the far end.
As illustrated below, this shouldn't be taken literally. Since the wires in this circuit are
pretty much perfect conductors, it really doesn't matter. As long as all the things that are
shown to be connected together are connected, the circuit will work.
Dual Lamp Circuit Wiring Alternatives

3-ELECTRICAL DEVICES

To move beyond cells and lamps, we need to introduce some more electrical and electronic devices and
their symbols.

Push button. A normally open push button conducts electricity when it is being pressed,
otherwise it's an open circuit.

Switch. Has an on and an off position. Conducts when it's on and is an open circuit when off.

To see how devices combine, the cell and lamp circuit from above is recreated below with the addition of a
switch to turn the lamp on and off. The switch works, just like it looks like in the diagram, by making or breaking
a connection which completes the circuit or leaves it open. An important observation is that it doesn't matter
whether the switch is on the connection from the positive side of the battery to the lamp or on the negative
side. As long as it can disrupt the circuit somewhere, it will work as a switch.

Lamp Circuit with switch

More sophisticated circuits require more complex components. Some more are presented below.

Resistor. Device that resists the flow of electricity equally in both directions. The two main
important values associated with resistors are their resistance and their power rating. Resistance is measured
in Ohms ( ). An Ohm is quite a small measurement for a lot of electrical applications so the k (or just k) is
often used. 1k is 1000 . The other value is power. Resistors dissipate energy so its important that exactly how
much energy they can dissipate is known. Most applications for resistors require only fractional Watts of power.

Capacitor. Device that temporarily stores electric charge. There are two main important values
that characterize a capacitor. The first is the capacitance - measured in Farads. It turns out that a Farad is a
huge amount, so capacitors are often measured in micro-Farads ( F) or pico-Farads (pF). The other important
quantity is the rated voltage. This value must never be exceeded in a circuit.

Diode. Semiconductor device that conducts electricity in only one direction. Exist in different
varieties. Zener diodes permit conduction in the reverse direction only when the reverse voltage exceeds a
certain amount. TVS diodes are like Zeners except capable of much higher currents.

MOSFET. (very short for metal-oxide silicon field-effect transistor). Special kind of transistor
switch. When the Gate (G) terminal voltage is brought sufficiently high, current will flow from the Drain (D) to
the Source (S) terminal. Usually require the Gate voltage to be 12V above the Source, but logic-level
MOSFET's can work directly from a microprocessor output. MOSFETs are characterized by several values:
their maximum voltage, their resistance when they're on and their maximum power dissipation. Be wary of
manufacturer's claims about maximum current often these are exaggerated and require the device be kept at
25°C which often requires liquid nitrogen cooling...

Potentiometer. A variable resistor. Often connected as a voltage divider to create variable


voltages when used as a rotational position sensor.

LED. Light Emitting Diode. Common indicator in electronics. Produces a lot of light for not
much current. But will very quickly (perhaps instantaneously) burn out if too much current is allowed to flow in
it. Like any diode, has very low resistance in its conducting direction, so a resistor in series with it to limit the
current is usually a requirement.

Photoresistor. A resistor with the useful property that its resistance changes depending on how
much light it is receiving. Photoresistors can have a quite impressive resistance range, for example from a few
million ohms (M ) in the dark to under a hundred ohms in bright light. One possible disadvantage is that their
reaction time is in the order of 100ms - too slow for many applications.

Supply. The Supply symbol is a diagrammatical shortcut used to indicate that the wire is
connected back to the power. It saves having to draw a wire from the power source to every point in the circuit
that uses it.

Ground. The Ground symbol is a diagrammatical shortcut of the same kind as the supply. It is
used to indicate that whatever is connected should be considered to be connected to the Ground of the power
supply.

Motor. Conventional DC motor. When deciding how to control a given motor there are several
important issues: what voltage was the motor designed to work with and how much current does it draw when
it's running. Commonly available DC motors can draw anything from 10mA to more than 100A. Motor selection
is a huge topic. You'll need to consider voltage, power consumption, RPM, torque, start and stall current,
mounting requirements, heat dissipation, etc.

Coil. Can represent a relay, a pneumatic or hydraulic valve or solenoid. The principle is the
same in all cases: when the coil is energized, it creates a magnetic field which attracts some metal part. Some
coils can heat up if left on for a long time so they are often given a duty-cycle meaning that their designer
specifies whether they can be left on indefinitely or whether they're designed only to switch on and off again
rarely.

Battery. The idea is that there is one wide and narrow line (cell) for each cell in the battery.
When it becomes onerous to draw all the cells, an ellipsis is added before the last cell. The common
rectangular 9V battery we buy at the store is in fact 6 1.5V cells stacked together.

4-OHM’S LAWS

We mentioned earlier that there is a relationship between voltage, current and resistance. It turns out
that this relationship is a mathematical one and it can be expressed very simply by way of Ohm's
Law.
Ohm's Law observes that in a simple resistive circuit, voltage (V) , resistance (R) and current (I) are
related in the following way:

V = IR

This expression can be rearranged algebraically to find other ways to use it as follows:

I = V / R (dividing both sides by R)


R = V / I (dividing both sides by I)

The idea is that if you know two of the quantities, you can work out the third by using one
of the equations.

Resistor

In the circuit fragment above, a resistor is connected between a 5V supply and ground (0V). We can use this to
check the assertions we made earlier about the effects of doubling the resistance on a circuit and so on. If the
resistor's value is 10 , we can work out what current will flow as follows. We know the voltage (5V) and the
resistance (10 ), so the form of the equation we need is:

I=V/R

Substituting our values in we have:

I = 5 / 10

The current I in our circuit will be 0.5A which can also be expressed as 500mA.

Now if we were to double the resistance (to 20 ), let's confirm that we halve the current:

I=V/R

Substituting our values in we have:

I = 5 / 20

The current I in our new circuit will be 0.25A (or 250mA), which is indeed half the previous current.

Ohm's Law is a very important relation to remember how to use. Finding current from resistance and voltage or
finding voltage from current and resistance is something that is required very frequently in electronics.

The energy of the electricity passing through the resistor is being converted by the resistor into waste heat.
The ability of the resistor to do this depends on how big it is. If it's too small it will not be able to get rid of the
heat fast enough and it may start to overheat and eventual burn.
The resistor's power rating is stated in Watts. Recall that the power is calculated by multiplying the voltage by
the current:

P = VI

In this last case, the voltage across the resistor was the same, 5V, but the current was halved to 0.25A. Putting
those values into the equation gives:

P = (5)(0.25)

P = 1.25W

Thus, the resistor chosen must be rated at least 1.25 W - which is rather a large resistor in the context of digital
electronics.

As a side note and another nice use of Ohm's Law, the V term in the power expression P = VI can be
substituted by IR, making P = I R, which a very convenient way to calculate power when the voltage drop
across something is not known.

5-VOLTAGE DIVIDERS

One of the most important concepts when interfacing sensors is the concept of the voltage divider.

Consider the circuit we were looking at in the last section. If we were to take a look at the
voltage at the high side of the resistor, we would measure approximately 5V. If we were to
measure the voltage at the low side we'd see approximately 0V. What would we measure in
the middle of the resistor? Presumably with half the resistive material above and half below
we'd measure 2.5V.

Voltage drop in resistor

If we have two resistors, this desire to peek inside a resistor can be faked. This stems from the observation
that when two resistors are connected in line (called in series) their resistances add linearly. Two 1000 (1k)
resistors in series have a total resistance of 2k, for example. Thus the point p between the two resistors
can be thought of as the midpoint of a 2k resistor.

Two Resistors

Here intuitively, it can be seen that there's a good case for suspecting that the voltage at point p is indeed 2.5V.
We can actually prove this with some simple math and using Ohm's Law. If math's not your strong point you
can skip lightly over the next paragraph or two.
So in our the circuit above, we have two resistors, of 1k each. Their total resistance is 2k, so, by Ohms law, the
current running through them is

I=V/R

I = 5 / 2000

I = 0.0025A or 2.5mA

Using Ohm's Law, we can actually work out what that voltage will be. The trick is that now we know the current
running through both of the resistors, we can apply Ohm's law to each individual one to see what the voltage
across it is. Taking the bottom resistor,

V = IR

V = 0.0025 * 1000

V = 2.5V

This means that at the point p there will be a voltage of 2.5V. So our intuition is correct.

Using two resistances (of potentially various kinds and values) a voltage somewhere between the supplied
voltage and zero can be obtained. If the high side resistor is converted into a smaller value than the low side
resistor, the voltage at p will be higher. If the high side resistor is converted into a higher value than the low
side resistor the voltage at p will be lower.

Now if you have a way of measuring voltages (like we do on a number of our Teleo modules) this can be a very
useful thing. We can see how useful in the next few example circuits.

Firstly there is the potentiometer. The potentiometer the device that used to control the volume and other
things on your stereo before everything became digital and microprocessor controlled. You turned it and things
got louder or quieter. The potentiometer is just a voltage divider. It has a resistive strip arranged in a circle
and an arm (called a wiper) that can tap onto it at any point on the circle. Power is applied to one end of the
resistive strip and ground to the other. The wiper can slide to any point in between.

will be at a value close to 5V. When


In the circuit below, when the wiper is close to the 5V the point p
it's halfway, it will be around 2.5V and when it's close to the other end it will be close to
0V.

Potentiometer

Another example of the usefulness of a voltage divider is in the use of a resistive sensor (like a photoresistor).
The circuit below shows how this is wired. This time point p measures 2.5V only if the resistance of the
photoresistor is almost exactly the same as the resistor below it (r). If the photoresistor's resistance falls (as
happens when it is exposed to light) the voltage at p will rise. Conversely if the photoresistor's resistance rises
(as happens when it is removed from the light), the voltage at p will fall. A device monitoring the voltage at line
p will therefore be able to monitor the amount of light on the photoresistor.

Light Sensor

On final example to illustrate the usefulness of the voltage divider is the push button used as a digital input to
electronics. The idea is to get a high voltage at p when the button is presses and a low one when it is
not.

Push Button without a Voltage Divider

The first question to ask is why bother with a voltage divider at all? Without it, won't the push button connect p
to 5V when it's pushed and leave it at 0V when un-pushed? This is true for the first part, but not true for the
second part. In fact what happens is that the button does indeed connect p to 5V when it's pushed, but when
it's released, the wire to p is actually unconnected (called floating). It is not getting pulled up to 5V, nor is it
getting pulled down to 0V. Electrical noise from nearby wires and other devices will make the actual voltage on
the wire fluctuate unpredictably. Sensitive electronics reading this voltage may get erroneous readings in this
state.

The solution is to make a voltage divider. When the switch is open, its resistance is very very high compared to
the resistor, so the voltage at p is very close to 0V. When the switch is closed, its resistance is
close to zero: very low when compared to the resistor, so the point p reads a voltage close
to 5V. In most modern digital circuits a fairly high value resistor (like 10k) can be used so
that the current being wasted by flowing through the switch and through the resistor is kept
to a fairly low amount. (this can be calculated using Ohm's Law: I = V / R, I = 5 / 10000, I
= 0.5mA).

Push ButtonVoltage Divider


When a resistor is used in this way - to make sure an input reads 0V when there is no circuitry driving the line
either way - it is called a pull-down resistor. A resistor tied to +5v, for example, to make sure an input reads a
positive voltage, when there is no driving circuitry is called a pull-up resistor.

6- DIGITAL LOGIC

Digital Logic is a whole subcategory of electronics stemming from the observation that by limiting signal values
to two discrete values (5V and 0V, say) you can represent binary values: i.e. values made up of only the digits
"0" and "1". Each one of these binary digits is called a bit (a contraction of "binary digit"). By convention, in a
5V system, 0V is assigned to the value 0 and 5V is assigned the value 1.

Strings of 0's and 1's can represent anything we like. We can have them represent numbers by using the
binary number system. The binary number system is exactly like our base 10 numbering system except there
are only the two numerals (0 and 1), not ten (0,1,2,3,4,5,6,7,8,9). In case you haven't done it for a while, in the
decimal system we count like this (although normally we don't put the leading zero's in).

000, 001, 002, 003, 004, 005, 006, 007, 008, 009, 010, 011, 012, 013, ... 095, 096, 097, 098, 099, 100, 101, ...

The idea is that as we count, we go through all of our numerals in the rightmost column, then when we run out
we increase the next column and go back to the 0 in the first column and repeat until the next column's
numerals are all used up and so on. In the binary system we count exactly the same way, except we only have
two numerals, so the next column thing happens a lot more often.

000, 001, 010, 011, 100, 101, 110, 111, ...


(0) (1) (2) (3) (4) (5) (6) (7)

You might think that this is kind of wasteful, since we're already using three digits (bits) by the time we've
counted to 4, but this is offset by the fact that each digit can be assigned to just one wire carrying a 0 or 1. With
three bits we can represent up to 8 things. With 8 bits (which is called a byte) we can represent up to 256
things. This number grows quickly, by the time you get to 64 bits, you can represent 18446744073709551616
things.

As you might suspect, you can add, subtract, multiply, divide, binary numbers very easily using the same
techniques you're familiar with in regular decimal arithmetic. You can even have binary decimal fractions
(111.01, for example is 7.25) and do all kinds of heavy mathematics. The things you can represent by binary
numbers can be bank balances, colors, calendar dates, distances, speeds, instructions - anything you like.
This is where computers come in - computers work with and by manipulation of binary numbers, all trading on
the idea that you can use wires carrying 0's and 1's to represent things in the real world and also to switch
circuitry on and off.

Microprocessors use 0's and 1's to do everything, so it should not be surprising that getting 0's and 1's in and
out of them is extremely easy. In a conventional 5V system, If an input pin is connected to 5V, it will read a "1".
If it's connected to a 0V it will read a "0". Similarly, if an output is set internally to 1 it will have 5V on the pin or
0V if a 0 set. Note that outputs can sometimes be disabled. When this happens the output is said to be
floating and could have any value.

A relatively new phenomenon is the presence of analog inputs on microprocessors. Analog inputs permit the
reading of voltages between 0V and the value of the chip's power supply voltage. (Note that a microprocessor
should never be directly connected to a voltage greater than it's power supply). Inside the microprocessor
equipped to handle analog input, there is a device called an analog to digital converter (ADC) which samples
the voltage on the pin and converts it into a binary number. The resolution of the ADC is just the number of
binary digits that it resolves the voltage into. If it were a 1bit ADC, it's one digit would be able to resolve the
voltage range into only two levels (i.e. < 2.5V and >2.5V). A 2bit ADC would resolve the voltage range into four
levels (00, 01, 10, 11), and so on. The PIC18 microprocessors have 10bit ADC's, so they can resolve the 0 -
5V range into 1024 levels - each one being 5V / 1024 = about 0.0049V.

7- PULSE WIDTH MODULATION (PWM)

It turns out that creating a variable voltage at more than a few hundred milliamps efficiently from a power
supply is not that easy. Pulse Width Modulation (PWM) is a common technique for using simple on-off
switching to produce the effect of a truly variable voltage. Motors, lights and a large variety of other devices
can be controlled in this way. House light dimmers work the same way.

What happens is that instead of a device either solidly on or off, it is switched on and off many times a second.
In the diagram below, voltage is plotted against time. The waveform in gray shows the rapid switching output of
the PWM circuit, and the effective voltage is shaded in as an overlay.

PWM Concept Diagram

On the left portion of the diagram, the power is only on for a small amount of time, so the effective voltage the
device under control experiences is very small. In the mid portion, the power is on for half of the total time, so
the controlled device experiences approximately half the full voltage. Finally in the right portion of the diagram,
the power is on for all the time, each cycle, in fact it never goes off at all, and so the voltage the device
experiences is the full value of the supply.

The portion of each cycle for which the power is switched on is called the duty cycle, and is often expressed
as a percentage ranging from 0% to 100%.

The effect this has on lights is like a dimmer - the light glows dimly and brightly depending on the duty cycle. If
this technique is applied to a normal DC brushed motor, the motor speeds up and slows down.

8-POWER SUPPLIES

There are many ways to provide power to a Teleo System. The best choice will depend on
many things: thecurrent requirements, portability, budget, etc. In all cases, the power supply
is presumed to have at least one output in the range 7.5V - 30V (the voltage range permitted
for Teleo Systems). The positive line should be connected to the positive (+) connector on
the Teleo Power Module. The ground (sometimes called "-") line should be connected to
the other connector.
Connection to the Power Module

Batteries provide a wide range of voltages and current capacities. They are portable and sometimes
rechargable. One of the major benefits of batteries (particulary NiCads and Lead Acid batteries) is their ability
to deliver huge amounts of current. Care should be take with car batteries and many battery packs not to short
circuit them. One obvious disadvantage is that they either need replacement or recharging before very long.

Linear Power Supplies used to be the main way to turn 110V AC (in the US) into DC power that most
electronic devices require. They are typically heavy for their current output, but units can be found to supply
huge amounts of current where this is necessary. The quality of the power obtained from a linear supply can
vary tremendously from perfect to only half rectified (meaning the voltage will vary from 0 volts to some higher
value 60 times a second). Unrectified power supplies have the additional problem that their voltage under light
loads may be significantly higher than under heavy loads. These have to be monitored closely so their output
voltage doesn't exceed the maximum permissible voltage.

Switch-mode Power Supplies can deliver a very high power output from a small package. They now form the
bulk of laptop and other power supply equipment. These power supplies are a little quirkier than their linear
cousins - sometimes generating electrical noise and sometimes in multiple output supplies requiring certain
minimum currents in order for them to work.

9- WIRING CONVENTIONS
Choosing a wiring color convention might seem unnecessary and may even run counter to the desired asthetic
effect for the machine under construction, but we recommend that you adopt a color convention when wiring. It
doesn't much matter what you choose, as long as you're consistent. Two that we try to always use are red for
positive and black for ground. You always know that a red line has power and shouldn't be connected to just
anywhere. We sometimes also adopt other conventions like green for inputs, orange for outputs, etc. The
general idea is to use the same color for all wires that do roughly the same thing. When it comes time to figure-
out what is wired to what these conventions will really help.

10- MULTIMETER USE

A multimeter (also known as a meter) is the first test instrument that we recommend you buy if you're
juststarting out in electronics. They range from US$20 up. Depending on which one you choose to get it will
have a different set of features, but most will at least allow you to measure voltage, current and resistance.
There are almost always at least two possible sockets for the positive (red) probe on the multimeter. One will
be for measuring voltage and resistance. The other(s) are for measuring current. It is important to put the
probe in the correct socket before using the multimeter.

Voltage measurement is achieved by making sure the probe is in the correct socket, selecting the voltage
setting, putting the negative (black) probe on the circuit's ground and the positive probe on the point you want
to test. You will get a reading of the voltage present. Some multimeters require that you pre-select a voltage
range through a variety of methods to enable it to more accurately provide a reading. In these meters, care
should be taken to not measure a large voltage when the meter is expecting a small one as damage to the
meter can result.

Resistance is measured when a device is not in a circuit or at least the circuit is not powered up. To prepare
the multimeter, select the resistance mode and make sure the positive probe is in the correct socket. In this
mode the multimeter puts out a small positive voltage on the positive probe and deduces from the current
being drawn, what the resistance must be using Ohm's Law . This mode is sometimes supplemented by a
continuity function where the meter beeps when the probes are connected via a very low resistance. This is
helpful when testing if two points on a device are connected as expected.

Current measurement is achieved by inserting the multimeter into a running circuit. To prepare the multimeter,
make sure the positive probe is in the correct socket (often marked A), The circuit actually needs to be broken
and re-routed through the multimeter. Be careful to make sure the multimeter can withstand the current you're
going to measure. Most have fuses, but it still pays to be careful since who wants to fool around changing
mutlimeter fuses all the time?

11-Wire Choice

Wire comes in an variety of shapes and forms. There are several key considerations when choosing wire:
current capacity (resistance) and mechanical strength being the primary ones that concern us here.

Above all, we recommend that you use stranded wire, rather than the solid wire. The stranded kind is much
more tolerant of mechanical movement. It will permit many more flexes before breaking than the solid kind.

Wire Gauges 30, 22, 18, 16 Next to a Quarter

In the US at least, wire is graded according to the American Wire Gauge (AWG) standard. You can see from
the photo above that different wire gauges are considerably different in size. We recommend the following:

AWG 30 - for wiring of very low current electronic signals over short distances where there is absolutely no
mechanical movement of the wire. Resistance is 4 /1000'

AWG 22 - is excellent for all kinds of hook-up wiring. Not too small to break all the time, and not too heavy to
restrict movement. Useful when the current doesn't exceed an amp or two. Very commonly used in long runs of
low current twisted pair cable. Resistance is 16.20 /1000'.

AWG 18 - quite heavy wire for high current use - up to 10A perhaps. Resistance is 6.3 /1000'
AWG 16 - very heavy duty wire for high current use - beyond 10A. Resistance is 4 /1000'

Adapted from: copyright © 2002-2004 MakingThings LLC


INSTITUTO UNIVERSITARIO DE TECNOLOGÍA
AGROINDUSTRIAL

OUTLINE

Prof. Elsa Rangel de Zambrano

SAN CRISTÒBAL, ______, _______, ______


ACTIVITY 4

De acuerdo a lo estudiado haga un esquema básico de la lectura siguiendo el orden


mostrado en el grammar 4.

_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________
_________________________________________________________________________

2 - Haga un resumen de la lectura anterior destacando las ideas mas


importantes desarrolladas por el autor en el texto.

FULLNAME:
_____________________________
_____________________________
_____________________________

______________________________ Score:_____________________
PASOS  PARA SEGUIR  EN  LA  LECTURA DE  UNA TABLA, CARTA  O
GRÁFICO
1. Lea  el título para determinar  lo que  abarca el  propósito del  gráfico o carta.
2. Lea  los  encabezados  sobre  la carta  o tabla  o  la  leyenda  sobre  el gráfico. La
leyenda  es  hecha  de  palabras  o  números  escritos  en los  lados  y el fondo  de  el
gráfico para decir  que  unidades de  medida  son  usadas.
3. Determine  cuales  cosas  están siendo  comparadas.
4. Estudie  el  gráfico o carta  notando  como  la relación  entre  2  o  más  cosas  esta
siendo  cambiada.
5.Trate de hacer una oración general explicando la tendencia o patrón que la
carta o el gráfico esta sugiriendo.

TIPOS DE GRÁFICOS
De Torta: Muestra como varias partes relacionan a un todo, con porcentajes.
Líneas: Indican una relación precisa entre 2 grupos de datos. Cada punto del
gráfico representa las 2 variables en relación a la otra. Es el tipo de gráfico
más confiable, muestra donde tiene lugar y las tendencias.
Barras: Permite la comparación de un pequeño número de valores ( poco mas de
10) tomados en diferentes tiempos. La presentación puede ser hecha
verticalmente u horizontalmente . Las barras pueden ser subdivididas en partes
de un todo o en porcentajes.
Sólidos o figuras: comparan 2 o mas totales usando figuras geométricas para
representar esas cantidades. Las figuras pueden ser cubos, esferas, cilindros.
Pictogramas: ilustran comparaciones aproximadas como lo hacen los gráficos de
barra pero usan figuras representativas como personas, vacas, autobuses u otros
artículos que estén siendo comparados.

COMO  LEER  LOS GRÁFICOS
1. Note  el  título y el  tipo de gráfico. Esto  indica  propósito,  idea  principal.
2. Note el arreglo de los datos. Lea los encabezados de las columnas ambos
vertical y horizontalmente para ver lo que está siendo comparado ( dólares, /
años,/ puntos, acres, etc.)
3. Note la escala ¿que es lo que aumenta el crecimiento y descenso?. Este alerta
a las alteraciones dentro del patrón que puedan cambiar la apariencia y causar
una mala interpretación.
4. Lea la leyenda. Esta indica el significado de los símbolos. Los códigos de
color y los patrones de la superficie (puntos donde se entrecruzan las
líneas ,etc.) son con frecuencia usados.
5. Note los símbolos dentro del gráfico. Esto puede ser solamente decorativos o
pueden ser componentes significativos.( consulte la leyenda).
6. Lea la información literal
7. Critique. Haga inferencias, saque conclusiones basados en los datos. Que
aplicaciones son posibles.
8. Relaciónelo con el material de texto.

Ejemplos:

The simulation plots both the input voltage (an AC signal of 1.5 volt peak amplitude and
2000 Hz frequency) and the current through the 15 volt battery, which is the same as the
current through the speaker. What we see here is a full AC sine wave alternating in both
positive and negative directions, and a half-wave output current waveform that only pulses
in one direction. If we were actually driving a speaker with this waveform, the sound
produced would be horribly distorted.
With the bias voltage source of 2.3 volts in place, the transistor remains in its active mode
throughout the entire cycle of the wave, faithfully reproducing the waveform at the speaker.
Notice that the input voltage (measured between nodes 1 and 0) fluctuates between about
0.8 volts and 3.8 volts, a peak-to-peak voltage of 3 volts just as expected (source voltage =
1.5 volts peak). The output (speaker) current varies between zero and almost 300 mA, 180 o
out of phase with the input (microphone) signal.
Plotted on the same scale (from 0 to 4 volts), we see that the output waveform ("+") has a
smaller peak-to-peak amplitude than the input waveform ("*"), in addition to being at a
lower bias voltage, not elevated up from 0 volts like the input. Since voltage gain for an AC
amplifier is defined by the ratio of AC amplitudes, we can ignore any DC bias separating
the two waveforms. Even so, the input waveform is still larger than the output, which tells
us that the voltage gain is less than 1 (a negative dB figure).

ACTIVE MODE OPERATION IN A TRANSISTOR

When a transistor is in the fully-off state (like an open switch), it is said to be cut
off. Conversely, when it is fully conductive between emitter and collector (passing
as much current through the collector as the collector power supply and load will
allow), it is said to be saturated. These are the two modes of operation explored
thus far in using the transistor as a switch.

However, bipolar transistors don't have to be restricted to these two extreme modes of
operation. As we learned in the previous section, base current "opens a gate" for a limited
amount of current through the collector. If this limit for the controlled current is greater
than zero but less than the maximum allowed by the power supply and load circuit, the
transistor will "throttle" the collector current in a mode somewhere between cut off and
saturation. This mode of operation is called the active mode.
An automotive analogy for transistor operation is as follows: cut off is the condition where
there is no motive force generated by the mechanical parts of the car to make it move. In
cut off mode, the brake is engaged (zero base current), preventing motion (collector
current). Active mode is when the automobile is cruising at a constant, controlled speed
(constant, controlled collector current) as dictated by the driver. Saturation is when the
automobile is driving up a steep hill that prevents it from going as fast as the driver would
wish. In other words, a "saturated" automobile is one where the accelerator pedal is pushed
all the way down (base current calling for more collector current than can be provided by
the power supply/load circuit).

I'll set up a circuit for SPICE simulation to demonstrate what happens when a transistor is
in its active mode of operation:

"Q" is the standard letter designation for a transistor in a schematic diagram, just as "R" is
for resistor and "C" is for capacitor. In this circuit, we have an NPN transistor powered by a
battery (V1) and controlled by current through a current source (I1). A current source is a
device that outputs a specific amount of current, generating as much or as little voltage as
necessary across its terminals to ensure that exact amount of current through it. Current
sources are notoriously difficult to find in nature (unlike voltage sources, which by contrast
attempt to maintain a constant voltage, outputting as much or as little current in the
fulfilment of that task), but can be simulated with a small collection of electronic
components. As we are about to see, transistors themselves tend to mimic the constant-
current behaviour of a current source in their ability to regulate current at a fixed value.

In the SPICE simulation, I'll set the current source at a constant value of 20 µA, then vary
the voltage source (V1) over a range of 0 to 2 volts and monitor how much current goes
through it. The "dummy" battery (V ammeter) with its output of 0 volts serves merely to
provide SPICE with a circuit element for current measurement.
bipolar transistor simulation
i1 0 1 dc 20u
q1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 npn
.dc v1 0 2 0.05
.plot dc i(vammeter)
.end

type npn
is 1.00E-16
bf 100.000
nf 1.000
br 1.000
nr 1.000

The constant base current of 20 µA sets a collector current limit of 2 mA, exactly 100 times
as much. Notice how flat the curve is for collector current over the range of battery voltage
from 0 to 2 volts. The only exception to this featureless plot is at the very beginning, where
the battery increases from 0 volts to 0.25 volts. There, the collector current increases
rapidly from 0 amps to its limit of 2 mA.
Let's see what happens if we vary the battery voltage over a wider range, this time from 0 to
50 volts. We'll keep the base current steady at 20 µA:

bipolar transistor simulation


i1 0 1 dc 20u
q1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 npn
.dc v1 0 50 2
.plot dc i(vammeter)
.end

type npn
is 1.00E-16
bf 100.000
nf 1.000
br 1.000
nr 1.000

Not surprisingly, SPICE gives us a similar plot: a flat line, holding steady this time at 7.5
mA -- exactly 100 times the base current -- over the range of battery voltages from just
above 0 volts to 50 volts. It appears that the base current is the deciding factor for collector
current, the V1 battery voltage being irrelevant so long as it's above a certain minimum
level.

This voltage/current relationship is entirely different from what we're used to seeing across
a resistor. With a resistor, current increases linearly as the voltage across it increases. Here,
with a transistor, current from emitter to collector stays limited at a fixed, maximum value
no matter how high the voltage across emitter and collector increases.

Often it is useful to superimpose several collector current/voltage graphs for different base
currents on the same graph. A collection of curves like this -- one curve plotted for each
distinct level of base current -- for a particular transistor is called the transistor's
characteristic curves:

Each curve on the graph reflects the collector current of the transistor, plotted over a range
of collector-to-emitter voltages, for a given amount of base current. Since a transistor tends
to act as a current regulator, limiting collector current to a proportion set by the base
current, it is useful to express this proportion as a standard transistor performance measure.
Specifically, the ratio of collector current to base current is known as the Beta ratio
(symbolized by the Greek letter β):
Sometimes the β ratio is designated as "hfe," a label used in a branch of mathematical
semiconductor analysis known as "hybrid parameters" which strives to achieve very precise
predictions of transistor performance with detailed equations. Hybrid parameter variables
are many, but they are all labeled with the general letter "h" and a specific subscript. The
variable "hfe" is just another (standardized) way of expressing the ratio of collector current
to base current, and is interchangeable with "β." Like all ratios, β is unitless.

β for any transistor is determined by its design: it cannot be altered after manufacture.
However, there are so many physical variables impacting β that it is rare to have two
transistors of the same design exactly match. If a circuit design relies on equal β ratios
between multiple transistors, "matched sets" of transistors may be purchased at extra cost.
However, it is generally considered bad design practice to engineer circuits with such
dependencies.

It would be nice if the β of a transistor remained stable for all operating conditions, but this
is not true in real life. For an actual transistor, the β ratio may vary by a factor of over 3
within its operating current limits. For example, a transistor with advertised β of 50 may
actually test with Ic/Ib ratios as low as 30 and as high as 100, depending on the amount of
collector current, the transistor's temperature, and frequency of amplified signal, among
other factors. For tutorial purposes it is adequate to assume a constant β for any given
transistor (which is what SPICE tends to do in a simulation), but just realize that real life is
not that simple!

Sometimes it is helpful for comprehension to "model" complex electronic components with


a collection of simpler, better-understood components. The following is a popular model
shown in many introductory electronics texts:
This model casts the transistor as a combination of diode and rheostat (variable resistor).
Current through the base-emitter diode controls the resistance of the collector-emitter
rheostat (as implied by the dashed line connecting the two components), thus controlling
collector current. An NPN transistor is modelled in the figure shown, but a PNP transistor
would be only slightly different (only the base-emitter diode would be reversed). This
model succeeds in illustrating the basic concept of transistor amplification: how the base
current signal can exert control over the collector current. However, I personally don't like
this model because it tends to miscommunicate the notion of a set amount of collector-
emitter resistance for a given amount of base current. If this were true, the transistor
wouldn't regulate collector current at all like the characteristic curves show. Instead of the
collector current curves flattening out after their brief rise as the collector-emitter voltage
increases, the collector current would be directly proportional to collector-emitter voltage,
rising steadily in a straight line on the graph.

A better transistor model, often seen in more advanced textbooks, is this:


It casts the transistor as a combination of diode and current source, the output of the current
source being set at a multiple (β ratio) of the base current. This model is far more accurate
in depicting the true input/output characteristics of a transistor: base current establishes a
certain amount of collector current, rather than a certain amount of collector-emitter
resistance as the first model implies. Also, this model is favoured when performing network
analysis on transistor circuits, the current source being a well-understood theoretical
component. Unfortunately, using a current source to model the transistor's current-
controlling behavior can be misleading: in no way will the transistor ever act as a source of
electrical energy, which the current source symbol implies is a possibility.

My own personal suggestion for a transistor model substitutes a constant-current diode for
the current source:
Since no diode ever acts as a source of electrical energy, this analogy escapes the false
implication of the current source model as a source of power, while depicting the
transistor's constant-current behaviour better than the rheostat model. Another way to
describe the constant-current diode's action would be to refer to it as a current regulator, so
this transistor illustration of mine might also be described as a diode-current regulator
model. The greatest disadvantage I see to this model is the relative obscurity of constant-
current diodes. Many people may be unfamiliar with their symbology or even of their
existence, unlike either rheostats or current sources, which are commonly known.

 REVIEW:
 A transistor is said to be in its active mode if it is operating somewhere between
fully on (saturated) and fully off (cut off).
 Base current tends to regulate collector current. By regulate, we mean that no more
collector current may exist than what is allowed by the base current.
 The ratio between collector current and base current is called "Beta" (β) or "hfe".
 β ratios are different for every transistor, and they tend to change for different
operating conditions.
Introduction

A transistor is a linear semiconductor device that controls current with the


application of a lower-power electrical signal. Transistors may be roughly
grouped into two major divisions: bipolar and field-effect. In the last chapter we
studied bipolar transistors, which utilize a small current to control a large
current. In this chapter, we'll introduce the general concept of the field-effect
transistor -- a device utilizing a small voltage to control current -- and then focus
on one particular type: the junction field-effect transistor. In the next chapter
we'll explore another type of field-effect transistor, the insulated gate variety.

All field-effect transistors are unipolar rather than bipolar devices. That is, the main current
through them is comprised either of electrons through an N-type semiconductor or holes
through a P-type semiconductor. This becomes more evident when a physical diagram of
the device is seen:

In a junction field-effect transistor, or JFET, the controlled current passes from source to
drain, or from drain to source as the case may be. The controlling voltage is applied
between the gate and source. Note how the current does not have to cross through a PN
junction on its way between source and drain: the path (called a channel) is an
uninterrupted block of semiconductor material. In the image just shown, this channel is an
N-type semiconductor. P-type channel JFETs are also manufactured:
Generally, N-channel JFETs are more commonly used than P-channel. The reasons for this
have to do with obscure details of semiconductor theory, which I'd rather not discuss in this
chapter. As with bipolar transistors, I believe the best way to introduce field-effect
transistor usage is to avoid theory whenever possible and concentrate instead on operational
characteristics. The only practical difference between N- and P-channel JFETs you need to
concern yourself with now is biasing of the PN junction formed between the gate material
and the channel.

With no voltage applied between gate and source, the channel is a wide-open path for
electrons to flow. However, if a voltage is applied between gate and source of such polarity
that it reverse-biases the PN junction, the flow between source and drain connections
becomes limited, or regulated, just as it was for bipolar transistors with a set amount of base
current. Maximum gate-source voltage "pinches off" all current through source and drain,
thus forcing the JFET into cutoff mode. This behavior is due to the depletion region of the
PN junction expanding under the influence of a reverse-bias voltage, eventually occupying
the entire width of the channel if the voltage is great enough. This action may be likened to
reducing the flow of a liquid through a flexible hose by squeezing it: with enough force, the
hose will be constricted enough to completely block the flow.
Note how this operational behavior is exactly opposite of the bipolar junction transistor.
Bipolar transistors are normally-off devices: no current through the base, no current through
the collector or the emitter. JFETs, on the other hand, are normally-on devices: no voltage
applied to the gate allows maximum current through the source and drain. Also take note
that the amount of current allowed through a JFET is determined by a voltage signal rather
than a current signal as with bipolar transistors. In fact, with the gate-source PN junction
reverse-biased, there should be nearly zero current through the gate connection. For this
reason, we classify the JFET as a voltage-controlled device, and the bipolar transistor as a
current-controlled device.

If the gate-source PN junction is forward-biased with a small voltage, the JFET channel
will "open" a little more to allow greater currents through. However, the PN junction of a
JFET is not built to handle any substantial current itself, and thus it is not recommended to
forward-bias the junction under any circumstances.

This is a very condensed overview of JFET operation. In the next section, we'll explore the
use of the JFET as a switching device.

The transistor as a switch

Like its bipolar cousin, the field-effect transistor may be used as an on/off switch
controlling electrical power to a load. Let's begin our investigation of the JFET as a switch
with our familiar switch/lamp circuit:
Remembering that the controlled current in a JFET flows between source and drain, we
substitute the source and drain connections of a JFET for the two ends of the switch in the
above circuit:

If you haven't noticed by now, the source and drain connections on a JFET look identical on
the schematic symbol. Unlike the bipolar junction transistor where the emitter is clearly
distinguished from the collector by the arrowhead, a JFET's source and drain lines both run
perpendicular into the bar representing the semiconductor channel. This is no accident, as
the source and drain lines of a JFET are often interchangeable in practice! In other words,
JFETs are usually able to handle channel current in either direction, from source to drain or
from drain to source.

Now all we need in the circuit is a way to control the JFET's conduction. With zero applied
voltage between gate and source, the JFET's channel will be "open," allowing full current to
the lamp. In order to turn the lamp off, we will need to connect another source of DC
voltage between the gate and source connections of the JFET like this:

Closing this switch will "pinch off" the JFET's channel, thus forcing it into cutoff and
turning the lamp off:
Note that there is no current going through the gate. As a reverse-biased PN junction, it
firmly opposes the flow of any electrons through it. As a voltage-controlled device, the
JFET requires negligible input current. This is an advantageous trait of the JFET over the
bipolar transistor: there is virtually zero power required of the controlling signal.

Opening the control switch again should disconnect the reverse-biasing DC voltage from
the gate, thus allowing the transistor to turn back on. Ideally, anyway, this is how it works.
In practice this may not work at all:

Why is this? Why doesn't the JFET's channel open up again and allow lamp current through
like it did before with no voltage applied between gate and source? The answer lies in the
operation of the reverse-biased gate-source junction. The depletion region within that
junction acts as an insulating barrier separating gate from source. As such, it possesses a
certain amount of capacitance capable of storing an electric charge potential. After this
junction has been forcibly reverse-biased by the application of an external voltage, it will
tend to hold that reverse-biasing voltage as a stored charge even after the source of that
voltage has been disconnected. What is needed to turn the JFET on again is to bleed off that
stored charge between the gate and source through a resistor:
This resistor's value is not very important. The capacitance of the JFET's gate-source
junction is very small, and so even a rather high-value bleed resistor creates a fast RC time
constant, allowing the transistor to resume conduction with little delay once the switch is
opened.

Like the bipolar transistor, it matters little where or what the controlling voltage comes
from. We could use a solar cell, thermocouple, or any other sort of voltage-generating
device to supply the voltage controlling the JFET's conduction. All that is required of a
voltage source for JFET switch operation is sufficient voltage to achieve pinch-off of the
JFET channel. This level is usually in the realm of a few volts DC, and is termed the pinch-
off or cutoff voltage. The exact pinch-off voltage for any given JFET is a function of its
unique design, and is not a universal figure like 0.7 volts is for a silicon BJT's base-emitter
junction voltage.

 REVIEW:
 Field-effect transistors control the current between source and drain connections by
a voltage applied between the gate and source. In a junction field-effect transistor
(JFET), there is a PN junction between the gate and source which is normally
reverse-biased for control of source-drain current.
 JFETs are normally-on (normally-saturated) devices. The application of a reverse-
biasing voltage between gate and source causes the depletion region of that junction
to expand, thereby "pinching off" the channel between source and drain through
which the controlled current travels.
 It may be necessary to attach a "bleed-off" resistor between gate and source to
discharge the stored charge built up across the junction's natural capacitance when
the controlling voltage is removed. Otherwise, a charge may remain to keep the
JFET in cutoff mode even after the voltage source has been disconnected.
Meter check of a transistor

Testing a JFET with a multimeter might seem to be a relatively easy task, seeing as how it
has only one PN junction to test: either measured between gate and source, or between gate
and drain.
Testing continuity through the drain-source channel is another matter, though. Remember
from the last section how a stored charge across the capacitance of the gate-channel PN
junction could hold the JFET in a pinched-off state without any external voltage being
applied across it? This can occur even when you're holding the JFET in your hand to test it!
Consequently, any meter reading of continuity through that channel will be unpredictable,
since you don't necessarily know if a charge is being stored by the gate-channel junction.
Of course, if you know beforehand which terminals on the device are the gate, source, and
drain, you may connect a jumper wire between gate and source to eliminate any stored
charge and then proceed to test source-drain continuity with no problem. However, if you
don't know which terminals are which, the unpredictability of the source-drain connection
may confuse your determination of terminal identity.

A good strategy to follow when testing a JFET is to insert the pins of the transistor into
anti-static foam (the material used to ship and store static-sensitive electronic components)
just prior to testing. The conductivity of the foam will make a resistive connection between
all terminals of the transistor when it is inserted. This connection will ensure that all
residual voltage built up across the gate-channel PN junction will be neutralized, thus
"opening up" the channel for an accurate meter test of source-to-drain continuity.

Since the JFET channel is a single, uninterrupted piece of semiconductor material, there is
usually no difference between the source and drain terminals. A resistance check from
source to drain should yield the same value as a check from drain to source. This resistance
should be relatively low (a few hundred ohms at most) when the gate-source PN junction
voltage is zero. By applying a reverse-bias voltage between gate and source, pinch-off of
the channel should be apparent by an increased resistance reading on the meter.

Active-mode operation

JFETs, like bipolar transistors, are able to "throttle" current in a mode between cutoff and
saturation called the active mode. To better understand JFET operation, let's set up a SPICE
simulation similar to the one used to explore basic bipolar transistor function:

jfet simulation
vin 0 1 dc 1
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 2 0.05
.plot dc i(vammeter)
.end

Note that the transistor labeled "Q1" in the schematic is represented in the SPICE netlist as
j1. Although all transistor types are commonly referred to as "Q" devices in circuit
schematics -- just as resistors are referred to by "R" designations, and capacitors by "C" --
SPICE needs to be told what type of transistor this is by means of a different letter
designation: q for bipolar junction transistors, and j for junction field-effect transistors.
Here, the controlling signal is a steady voltage of 1 volt, applied with negative towards the
JFET gate and positive toward the JFET source, to reverse-bias the PN junction. In the first
BJT simulation of chapter 4, a constant-current source of 20 µA was used for the
controlling signal, but remember that a JFET is a voltage-controlled device, not a current-
controlled device like the bipolar junction transistor.

Like the BJT, the JFET tends to regulate the controlled current at a fixed level above a
certain power supply voltage, no matter how high that voltage may climb. Of course, this
current regulation has limits in real life -- no transistor can withstand infinite voltage from a
power source -- and with enough drain-to-source voltage the transistor will "break down"
and drain current will surge. But within normal operating limits the JFET keeps the drain
current at a steady level independent of power supply voltage. To verify this, we'll run
another computer simulation, this time sweeping the power supply voltage (V1) all the way
to 50 volts:

jfet simulation
vin 0 1 dc 1
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end
Sure enough, the drain current remains steady at a value of 100 µA (1.000E-04 amps) no
matter how high the power supply voltage is adjusted.

Because the input voltage has control over the constriction of the JFET's channel, it makes
sense that changing this voltage should be the only action capable of altering the current
regulation point for the JFET, just like changing the base current on a BJT is the only action
capable of altering collector current regulation. Let's decrease the input voltage from 1 volt
to 0.5 volts and see what happens:

jfet simulation
vin 0 1 dc 0.5
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end
As expected, the drain current is greater now than it was in the previous simulation. With
less reverse-bias voltage impressed across the gate-source junction, the depletion region is
not as wide as it was before, thus "opening" the channel for charge carriers and increasing
the drain current figure.

Please note, however, the actual value of this new current figure: 225 µA (2.250E-04
amps). The last simulation showed a drain current of 100 µA, and that was with a gate-
source voltage of 1 volt. Now that we've reduced the controlling voltage by a factor of 2
(from 1 volt down to 0.5 volts), the drain current increased, but not by the same 2:1
proportion! Let's reduce our gate-source voltage once more by another factor of 2 (down to
0.25 volts) and see what happens:

jfet simulation
vin 0 1 dc 0.25
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end
With the gate-source voltage set to 0.25 volts, one-half what it was before, the drain current
is 306.3 µA. Although this is still an increase over the 225 µA from the prior simulation, it
isn't proportional to the change of the controlling voltage.

To obtain a better understanding of what is going on here, we should run a different kind of
simulation: one that keeps the power supply voltage constant and instead varies the
controlling (voltage) signal. When this kind of simulation was run on a BJT, the result was
a straight-line graph, showing how the input current / output current relationship of a BJT is
linear. Let's see what kind of relationship a JFET exhibits:

jfet simulation
vin 0 1 dc
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc 25
.model mod1 njf
.dc vin 0 2 0.1
.plot dc i(vammeter)
.end
This simulation directly reveals an important characteristic of the junction field-effect
transistor: the control effect of gate voltage over drain current is nonlinear. Notice how the
drain current does not decrease linearly as the gate-source voltage is increased. With the
bipolar junction transistor, collector current was directly proportional to base current:
output signal proportionately followed input signal. Not so with the JFET! The controlling
signal (gate-source voltage) has less and less effect over the drain current as it approaches
cutoff. In this simulation, most of the controlling action (75 percent of drain current
decrease -- from 400 µA to 100 µA) takes place within the first volt of gate-source voltage
(from 0 to 1 volt), while the remaining 25 percent of drain current reduction takes another
whole volt worth of input signal. Cutoff occurs at 2 volts input.

Linearity is generally important for a transistor because it allows it to faithfully amplify a


waveform without distorting it. If a transistor is nonlinear in its input/output amplification,
the shape of the input waveform will become corrupted in some way, leading to the
production of harmonics in the output signal. The only time linearity is not important in a
transistor circuit is when it's being operated at the extreme limits of cutoff and saturation
(off and on, respectively, like a switch).

A JFET's characteristic curves display the same current-regulating behavior as for a BJT,
and the nonlinearity between gate-to-source voltage and drain current is evident in the
disproportionate vertical spacings between the curves:
To better comprehend the current-regulating behavior of the JFET, it might be helpful to
draw a model made up of simpler, more common components, just as we did for the BJT:
In the case of the JFET, it is the voltage across the reverse-biased gate-source diode which
sets the current regulation point for the pair of constant-current diodes. A pair of opposing
constant-current diodes is included in the model to facilitate current in either direction
between source and drain, a trait made possible by the unipolar nature of the channel. With
no PN junctions for the source-drain current to traverse, there is no polarity sensitivity in
the controlled current. For this reason, JFETs are often referred to as bilateral devices.

A contrast of the JFET's characteristic curves against the curves for a bipolar transistor
reveals a notable difference: the linear (straight) portion of each curve's non-horizontal area
is surprisingly long compared to the respective portions of a BJT's characteristic curves:
A JFET transistor operated in the triode region tends to act very much like a plain resistor
as measured from drain to source. Like all simple resistances, its current/voltage graph is a
straight line. For this reason, the triode region (non-horizontal) portion of a JFET's
characteristic curve is sometimes referred to as the ohmic region. In this mode of operation
where there isn't enough drain-to-source voltage to bring drain current up to the regulated
point, the drain current is directly proportional to the drain-to-source voltage. In a carefully
designed circuit, this phenomenon can be used to an advantage. Operated in this region of
the curve, the JFET acts like a voltage-controlled resistance rather than a voltage-controlled
current regulator, and the appropriate model for the transistor is different:

Here and here alone the rheostat (variable resistor) model of a transistor is accurate. It must
be remembered, however, that this model of the transistor holds true only for a narrow
range of its operation: when it is extremely saturated (far less voltage applied between drain
and source than what is needed to achieve full regulated current through the drain). The
amount of resistance (measured in ohms) between drain and source in this mode is
controlled by how much reverse-bias voltage is applied between gate and source. The less
gate-to-source voltage, the less resistance (steeper line on graph).

Because JFETs are voltage-controlled current regulators (at least when they're allowed to
operate in their active), their inherent amplification factor cannot be expressed as a unitless
ratio as with BJTs. In other words, there is no β ratio for a JFET. This is true for all voltage-
controlled active devices, including other types of field-effect transistors and even electron
tubes. There is, however, an expression of controlled (drain) current to controlling (gate-
source) voltage, and it is called transconductance. Its unit is Siemens, the same unit for
conductance (formerly known as the mho).
Why this choice of units? Because the equation takes on the general form of current (output
signal) divided by voltage (input signal).

Unfortunately, the transconductance value for any JFET is not a stable quantity: it varies
significantly with the amount of gate-to-source control voltage applied to the transistor. As
we saw in the SPICE simulations, the drain current does not change proportionally with
changes in gate-source voltage. To calculate drain current for any given gate-source
voltage, there is another equation that may be used. It is obviously nonlinear upon
inspection (note the power of 2), reflecting the nonlinear behavior we've already
experienced in simulation:

 REVIEW:
 In their active modes, JFETs regulate drain current according to the amount of
reverse-bias voltage applied between gate and source, much like a BJT regulates
collector current according to base current. The mathematical ratio between drain
current (output) and gate-to-source voltage (input) is called transconductance, and it
is measured in units of Siemens.
 The relationship between gate-source (control) voltage and drain (controlled)
current is nonlinear: as gate-source voltage is decreased, drain current increases
exponentially. That is to say, the transconductance of a JFET is not constant over its
range of operation.
 In their triode region, JFETs regulate drain-to-source resistance according to the
amount of reverse-bias voltage applied between gate and source. In other words,
they act like voltage-controlled resistances.
Hysteresis

Thyristors are a class of semiconductor components exhibiting hysteresis, that


property whereby a system fails to return to its original state after some cause of state
change has been removed. A very simple example of hysteresis is the mechanical
action of a toggle switch: when the lever is pushed, it flips to one of two extreme
states (positions) and will remain there even after the source of motion is removed
(after you remove your hand from the switch lever). To illustrate the absence of
hysteresis, consider the action of a "momentary" pushbutton switch, which returns to
its original state after the button is no longer pressed: when the stimulus is removed
(your hand), the system (switch) immediately and fully returns to its prior state with
no "latching" behavior.

Bipolar, junction field-effect, and insulated gate field-effect transistors are all non-
hysteretic devices. That is, they do not inherently "latch" into a state after being stimulated
by a voltage or current signal. For any given input signal at any given time, a transistor will
exhibit a predictable output response as defined by its characteristic curve. Thyristors, on
the other hand, are semiconductor devices that tend to stay "on" once turned on, and tend to
stay "off" once turned off. A momentary event is able to flip these devices into either their
on or off states where they will remain that way on their own, even after the cause of the
state change is taken away. As such, they are useful only as on/off switching devices --
much like a toggle switch -- and cannot be used as analog signal amplifiers.
Thyristors are constructed using the same technology as bipolar junction transistors, and in
fact may be analyzed as circuits comprised of transistor pairs. How then, can a hysteretic
device (a thyristor) be made from non-hysteretic devices (transistors)? The answer to this
question is positive feedback, also known as regenerative feedback. As you should recall,
feedback is the condition where a percentage of the output signal is "fed back" to the input
of an amplifying device. Negative, or degenerative, feedback results in a diminishing of
voltage gain with increases in stability, linearity, and bandwidth. Positive feedback, on the
other hand, results in a kind of instability where the amplifier's output tends to "saturate." In
the case of thyristors, this saturating tendency equates to the device "wanting" to stay on
once turned on, and off once turned off.

In this chapter we will explore several different kinds of thyristors, most of which stem
from a single, basic two-transistor core circuit. Before we do that, though, it would be
beneficial to study the technological predecessor to thyristors: gas discharge tubes.

Gas discharge tubes

If you've ever witnessed a lightning storm, you've seen electrical hysteresis in action (and
probably didn't realize what you were seeing). The action of strong wind and rain
accumulates tremendous static electric charges between cloud and earth, and between
clouds as well. Electric charge imbalances manifest themselves as high voltages, and when
the electrical resistance of air can no longer hold these high voltages at bay, huge surges of
current travel between opposing poles of electrical charge which we call "lightning."

The buildup of high voltages by wind and rain is a fairly continuous process, the rate of
charge accumulation increasing under the proper atmospheric conditions. However,
lightning bolts are anything but continuous: they exist as relatively brief surges rather than
continuous discharges. Why is this? Why don't we see soft, glowing lightning arcs instead
of violently brief lightning bolts? The answer lies in the nonlinear (and hysteretic)
resistance of air.

Under ordinary conditions, air has an extremely high amount of resistance. It is so high, in
fact, that we typically treat its resistance as infinite and electrical conduction through the air
as negligible. The presence of water and/or dust in air lowers its resistance some, but it is
still an insulator for most practical purposes. When a sufficient amount of high voltage is
applied across a distance of air, though, its electrical properties change: electrons become
"stripped" from their normal positions around their respective atoms and are liberated to
constitute a current. In this state, air is considered to be ionized and is referred to as a
plasma rather than a normal gas. This usage of the word "plasma" is not to be confused
with the medical term (meaning the fluid portion of blood), but is a fourth state of matter,
the other three being solid, liquid, and vapor (gas). Plasma is a relatively good conductor of
electricity, its specific resistance being much lower than that of the same substance in its
gaseous state.

As an electric current moves through the plasma, there is energy dissipated in the plasma in
the form of heat, just as current through a solid resistor dissipates energy in the form of
heat. In the case of lightning, the temperatures involved are extremely high. High
temperatures are also sufficient to convert gaseous air into a plasma or maintain plasma in
that state without the presence of high voltage. As the voltage between cloud and earth, or
between cloud and cloud, decreases as the charge imbalance is neutralized by the current of
the lightning bolt, the heat dissipated by the bolt maintains the air path in a plasma state,
keeping its resistance low. The lightning bolt remains a plasma until the voltage decreases
to too low a level to sustain enough current to dissipate enough heat. Finally, the air returns
to a normal, gaseous state and stops conducting current, thus allowing voltage to build up
once more.

Note how throughout this cycle, the air exhibits hysteresis. When not conducting electricity,
it tends to remain an insulator until voltage builds up past a critical threshold point. Then,
once it changes state and becomes a plasma, it tends to remain a conductor until voltage
falls below a lower critical threshold point. Once "turned on" it tends to stay "on," and once
"turned off" it tends to stay "off." This hysteresis, combined with a steady buildup of
voltage due to the electrostatic effects of wind and rain, explains the action of lightning as
brief bursts.

In electronic terms, what we have here in the action of lightning is a simple relaxation
oscillator. Oscillators are electronic circuits that produce an oscillating (AC) voltage from a
steady supply of DC power. A relaxation oscillator is one that works on the principle of a
charging capacitor that is suddenly discharged every time its voltage reaches a critical
threshold value. One of the simplest relaxation oscillators in existence is comprised of three
components (not counting the DC power supply): a resistor, capacitor, and neon lamp:

Neon lamps are nothing more than two metal electrodes inside a sealed glass bulb,
separated by the neon gas inside. At room temperatures and with no applied voltage, the
lamp has nearly infinite resistance. However, once a certain threshold voltage is exceeded
(this voltage depends on the gas pressure and geometry of the lamp), the neon gas will
become ionized (turned into a plasma) and its resistance dramatically reduced. In effect, the
neon lamp exhibits the same characteristics as air in a lightning storm, complete with the
emission of light as a result of the discharge, albeit on a much smaller scale.
The capacitor in the relaxation oscillator circuit shown above charges at an inverse
exponential rate determined by the size of the resistor. When its voltage reaches the
threshold voltage of the lamp, the lamp suddenly "turns on" and quickly discharges the
capacitor to a low voltage value. Once discharged, the lamp "turns off" and allows the
capacitor to build up a charge once more. The result is a series of brief flashes of light from
the lamp, the rate of which dictated by battery voltage, resistor resistance, capacitor
capacitance, and lamp threshold voltage.

While gas-discharge lamps are more commonly used as sources of illumination, their
hysteretic properties were leveraged in slightly more sophisticated variants known as
thyratron tubes. Essentially a gas-filled triode tube (a triode being a three-element vacuum
electron tube performing much a similar function to the N-channel, D-type IGFET), the
thyratron tube could be turned on with a small control voltage applied between grid and
cathode, and turned off by reducing the plate-to-cathode voltage.

In essence, thyratron tubes were controlled versions of neon lamps built specifically for
switching current to a load. The dot inside the circle of the schematic symbol indicates a
gas fill, as opposed to the hard vacuum normally seen in other electron tube designs. In the
circuit shown above, the thyratron tube allows current through the load in one direction
(note the polarity across the load resistor) when triggered by the small DC control voltage
connected between grid and cathode. Note that the load's power source is AC, which
provides a clue as to how the thyratron turns off after it's been triggered on: since AC
voltage periodically passes through a condition of 0 volts between half-cycles, the current
through an AC-powered load must also periodically halt. This brief pause of current
between half-cycles gives the tube's gas time to cool, letting it return to its normal "off"
state. Conduction may resume only if there is enough voltage applied by the AC power
source (some other time in the wave's cycle) and if the DC control voltage allows it.

An oscilloscope display of load voltage in such a circuit would look something like this:
As the AC supply voltage climbs from zero volts to its first peak, the load voltage remains
at zero (no load current) until the threshold voltage is reached. At that point, the tube
switches "on" and begins to conduct, the load voltage now following the AC voltage
through the rest of the half cycle. Notice how there is load voltage (and thus load current)
even when the AC voltage waveform has dropped below the threshold value of the tube.
This is hysteresis at work: the tube stays in its conductive mode past the point where it first
turned on, continuing to conduct until there the supply voltage drops off to almost zero
volts. Because thyratron tubes are one-way (diode) devices, there is no voltage across the
load through the negative half-cycle of AC. In practical thyratron circuits, multiple tubes
arranged in some form of full-wave rectifier circuit to facilitate full-wave DC power to the
load.

Although I'm not sure if this was ever done, someone could have applied the thyratron tube
to a relaxation oscillator circuit and control the frequency with a small DC voltage between
grid and cathode, making a crude voltage-controlled oscillator, otherwise known as a VCO.
Relaxation oscillators tend to have poor frequency control, not to mention a very non-
sinusoidal output, and so they exist mostly as demonstration circuits (as is the case here) or
in applications where precise frequency control isn't important. Consequently, this use of a
thyratron tube would not have been a very practical one.
I speak of thyratron tubes in the past tense for good reason: modern semiconductor
components have obsoleted thyratron tube technology for all but a few very special
applications. It is no coincidence that the word thyristor bears so much similarity to the
word thyratron, for this class of semiconductor components does much the same thing: use
hysteretically switch current on and off. It is these modern devices that we now turn our
attention to.

 REVIEW:
 Electrical hysteresis, the tendency for a component to remain "on" (conducting)
after it begins to conduct and to remain "off" (nonconducting) after it ceases to
conduct, helps to explain why lightning bolts exist as momentary surges of current
rather than continuous discharges through the air.
 Simple gas-discharge tubes such as neon lamps exhibit electrical hysteresis.
 More advanced gas-discharge tubes have been made with control elements so that
their "turn-on" voltage could be adjusted by an external signal. The most common
of these tubes was called the thyratron.
 Simple oscillator circuits called relaxation oscillators may be created with nothing
more than a resistor-capacitor charging network and a hysteretic device connected
across the capacitor.
The Shockley Diode

Our exploration of thyristors begins with a device called the four-layer diode, also known
as a PNPN diode, or a Shockley diode after its inventor, William Shockley. This is not to be
confused with a Schottky diode, that two-layer metal-semiconductor device known for its
high switching speed. A crude illustration of the Shockley diode, often seen in textbooks, is
a four-layer sandwich of P-N-P-N semiconductor material:

Unfortunately, this simple illustration does nothing to enlighten the viewer on how it works
or why. Consider an alternative rendering of the device's construction:

Shown like this, it appears to be a set of interconnected bipolar transistors, one PNP and the
other NPN. Drawn using standard schematic symbols, and respecting the layer doping
concentrations not shown in the last image, the Shockley diode looks like this:
Let's connect one of these devices to a source of variable voltage and see what happens:

With no voltage applied, of course there will be no current. As voltage is initially increased,
there will still be no current because neither transistor is able to turn on: both will be in
cutoff mode. To understand why this is, consider what it takes to turn a bipolar junction
transistor on: current through the base-emitter junction. As you can see in the diagram, base
current through the lower transistor is controlled by the upper transistor, and the base
current through the upper transistor is controlled by the lower transistor. In other words,
neither transistor can turn on until the other transistor turns on. What we have here, in
vernacular terms, is known as a Catch-22.

So how can a Shockley diode ever conduct current, if its constituent transistors stubbornly
maintain themselves in a state of cutoff? The answer lies in the behavior of real transistors
as opposed to ideal transistors. An ideal bipolar transistor will never conduct collector
current if there is no base current, no matter how much or little voltage we apply between
collector and emitter. Real transistors, on the other hand, have definite limits to how much
collector-emitter voltage they can withstand before they break down and conduct. If two
real transistors are connected together in this fashion to form a Shockley diode, they will be
able to conduct if there is sufficient voltage applied by the battery between anode and
cathode to cause one of them to break down. Once one transistor breaks down and begins to
conduct, it will allow base current through the other transistor, causing it to turn on in a
normal fashion, which then allows base current through the first transistor. The end result is
that both transistors will be saturated, now keeping each other turned on instead of off.
So, we can force a Shockley diode to turn on by applying sufficient voltage between anode
and cathode. As we have seen, this will inevitably cause one of the transistors to turn on,
which then turns the other transistor on, ultimately "latching" both transistors on where they
will tend to remain. But how do we now get the two transistors to turn off again? Even if
the applied voltage is reduced to a point well below what it took to get the Shockley diode
conducting, it will remain conducting because both transistors now have base current to
maintain regular, controlled conduction. The answer to this is to reduce the applied voltage
to a much lower point where there is too little current to maintain transistor bias, at which
point one of the transistors will cutoff, which then halts base current through the other
transistor, sealing both transistors in the "off" state as they were before any voltage was
applied at all.

If we graph this sequence of events and plot the results on an I/V graph, the hysteresis is
very evident. First, we will observe the circuit as the DC voltage source (battery) is set to
zero voltage:

Next, we will steadily increase the DC voltage. Current through the circuit is at or nearly at
zero, as the breakdown limit has not been reached for either transistor:
When the voltage breakdown limit of one transistor is reached, it will begin to conduct
collector current even though no base current has gone through it yet. Normally, this sort of
treatment would destroy a bipolar junction transistor, but the PNP junctions comprising a
Shockley diode are engineered to take this kind of abuse, similar to the way a Zener diode
is built to handle reverse breakdown without sustaining damage. For the sake of illustration
I'll assume the lower transistor breaks down first, sending current through the base of the
upper transistor:

As the upper transistor receives base current, it turns on as expected. This action allows the
lower transistor to conduct normally, the two transistors "sealing" themselves in the "on"
state. Full current is very quickly seen in the circuit:
The positive feedback mentioned earlier in this chapter is clearly evident here. When one
transistor breaks down, it allows current through the device structure. This current may be
viewed as the "output" signal of the device. Once an output current is established, it works
to hold both transistors in saturation, thus ensuring the continuation of a substantial output
current. In other words, an output current "feeds back" positively to the input (transistor
base current) to keep both transistors in the "on" state, thus reinforcing (or regenerating)
itself.

With both transistors maintained in a state of saturation with the presence of ample base
current, they will continue to conduct even if the applied voltage is greatly reduced from
the breakdown level. The effect of positive feedback is to keep both transistors in a state of
saturation despite the loss of input stimulus (the original, high voltage needed to break
down one transistor and cause a base current through the other transistor):

If the DC voltage source is turned down too far, though, the circuit will eventually reach a
point where there isn't enough current to sustain both transistors in saturation. As one
transistor passes less and less collector current, it reduces the base current for the other
transistor, thus reducing base current for the first transistor. The vicious cycle continues
rapidly until both transistors fall into cutoff:

Here, positive feedback is again at work: the fact that the cause/effect cycle between both
transistors is "vicious" (a decrease in current through one works to decrease current through
the other, further decreasing current through the first transistor) indicates a positive
relationship between output (controlled current) and input (controlling current through the
transistors' bases).

The resulting curve on the graph is classicly hysteretic: as the input signal (voltage) is
increased and decreased, the output (current) does not follow the same path going down as
it did going up:

Put in simple terms, the Shockley diode tends to stay on once it's turned on, and stay off
once it's turned off. There is no "in-between" or "active" mode in its operation: it is a purely
on or off device, as are all thyristors.

There are a few special terms applied to Shockley diodes and all other thyristor devices
built upon the Shockley diode foundation. First is the term used to describe its "on" state:
latched. The word "latch" is reminiscent of a door lock mechanism, which tends to keep the
door closed once it has been pushed shut. The term firing refers to the initiation of a latched
state. In order to get a Shockley diode to latch, the applied voltage must be increased until
breakover is attained. Despite the fact that this action is best described in terms of transistor
breakdown, the term breakover is used instead because the end result is a pair of transistors
in mutual saturation rather than destruction as would be the case with a normal transistor. A
latched Shockley diode is re-set back into its nonconducting state by reducing current
through it until low-current dropout occurs.

It should be noted that Shockley diodes may be fired in a way other than breakover:
excessive voltage rise, or dv/dt. This is when the applied voltage across the diode increases
at a high rate of change. This is able to cause latching (turning on) of the diode due to
inherent junction capacitances within the transistors. Capacitors, as you may recall, oppose
changes in voltage by drawing or supplying current. If the applied voltage across a
Shockley diode rises at too fast a rate, those tiny capacitances will draw enough current
during that time to activate the transistor pair, turning them both on. Usually, this form of
latching is undesirable, and can be minimized by filtering high-frequency (fast voltage
rises) from the diode with series inductors and/or parallel resistor-capacitor networks called
snubbers:

The voltage rise limit of a Shockley diode is referred to as the critical rate of voltage rise.
Manufacturers usually provide this specification for the devices they sell.

 REVIEW:
 Shockley diodes are four-layer PNPN semiconductor devices. They behave as a pair
of interconnected PNP and NPN transistors.
 Like all thyristors, Shockley diodes tend to stay on once they've been turned on
(latched), and stay off once they've been turned off.
 There are two ways to latch a Shockley diode: exceed the anode-to-cathode
breakover voltage, or exceed the anode-to-cathode critical rate of voltage rise.
 There is only one way to cause a Shockley diode to stop conducting, and that is to
reduce the current going through it to a level below its low-current dropout
threshold.
The DIAC

Like all diodes, Shockley diodes are unidirectional devices; that is, they only conduct
current in one direction. If bidirectional (AC) operation is desired, two Shockley diodes
may be joined in parallel facing different directions to form a new kind of thyristor, the
DIAC:

A DIAC operated with a DC voltage across it behaves exactly the same as a Shockley
diode. With AC, however, the behavior is different from what one might expect. Because
alternating current repeatedly reverses direction, DIACs will not stay latched longer than
one-half cycle. If a DIAC becomes latched, it will continue to conduct current only as long
as there is voltage available to push enough current in that direction. When the AC polarity
reverses, as it must twice per cycle, the DIAC will drop out due to insufficient current,
necessitating another breakover before it conducts again. The result is a current waveform
that looks like this:

DIACs are almost never used alone, but in conjunction with other thyristor devices.
The Silicon-Controlled Rectifier (SCR)

Shockley diodes are curious devices, but rather limited in application. Their usefulness may
be expanded, however, by equipping them with another means of latching. In doing so, they
become true amplifying devices (if only in an on/off mode), and we refer to them as
silicon-controlled rectifiers, or SCRs.

The progression from Shockley diode to SCR is achieved with one small addition, actually
nothing more than a third wire connection to the existing PNPN structure:

If an SCR's gate is left floating (disconnected), it behaves exactly as a Shockley diode. It


may be latched by breakover voltage or by exceeding the critical rate of voltage rise
between anode and cathode, just as with the Shockley diode. Dropout is accomplished by
reducing current until one or both internal transistors fall into cutoff mode, also like the
Shockley diode. However, because the gate terminal connects directly to the base of the
lower transistor, it may be used as an alternative means to latch the SCR. By applying a
small voltage between gate and cathode, the lower transistor will be forced on by the
resulting base current, which will cause the upper transistor to conduct, which then supplies
the lower transistor's base with current so that it no longer needs to be activated by a gate
voltage. The necessary gate current to initiate latch-up, of course, will be much lower than
the current through the SCR from cathode to anode, so the SCR does achieve a measure of
amplification.

This method of securing SCR conduction is called triggering, and it is by far the most
common way that SCRs are latched in actual practice. In fact, SCRs are usually chosen so
that their breakover voltage is far beyond the greatest voltage expected to be experienced
from the power source, so that it can be turned on only by an intentional voltage pulse
applied to the gate.

It should be mentioned that SCRs may sometimes be turned off by directly shorting their
gate and cathode terminals together, or by "reverse-triggering" the gate with a negative
voltage (in reference to the cathode), so that the lower transistor is forced into cutoff. I say
this is "sometimes" possible because it involves shunting all of the upper transistor's
collector current past the lower transistor's base. This current may be substantial, making
triggered shut-off of an SCR difficult at best. A variation of the SCR, called a Gate-Turn-
Off thyristor, or GTO, makes this task easier. But even with a GTO, the gate current
required to turn it off may be as much as 20% of the anode (load) current! The schematic
symbol for a GTO is shown in the following illustration:

SCRs and GTOs share the same equivalent schematics (two transistors connected in a
positive-feedback fashion), the only differences being details of construction designed to
grant the NPN transistor a greater β than the PNP. This allows a smaller gate current
(forward or reverse) to exert a greater degree of control over conduction from cathode to
anode, with the PNP transistor's latched state being more dependent upon the NPN's than
vice versa. The Gate-Turn-Off thyristor is also known by the name of Gate-Controlled
Switch, or GCS.

A rudimentary test of SCR function, or at least terminal identification, may be performed


with an ohmmeter. Because the internal connection between gate and cathode is a single PN
junction, a meter should indicate continuity between these terminals with the red test lead
on the gate and the black test lead on the cathode like this:
All other continuity measurements performed on an SCR will show "open" ("OL" on some
digital multimeter displays). It must be understood that this test is very crude and does not
constitute a comprehensive assessment of the SCR. It is possible for an SCR to give good
ohmmeter indications and still be defective. Ultimately, the only way to test an SCR is to
subject it to a load current.

If you are using a multimeter with a "diode check" function, the gate-to-cathode junction
voltage indication you get may or may not correspond to what's expected of a silicon PN
junction (approximately 0.7 volts). In some cases, you will read a much lower junction
voltage: mere hundredths of a volt. This is due to an internal resistor connected between the
gate and cathode incorporated within some SCRs. This resistor is added to make the SCR
less susceptible to false triggering by spurious voltage spikes, from circuit "noise" or from
static electric discharge. In other words, having a resistor connected across the gate-cathode
junction requires that a strong triggering signal (substantial current) be applied to latch the
SCR. This feature is often found in larger SCRs, not on small SCRs. Bear in mind that an
SCR with an internal resistor connected between gate and cathode will indicate continuity
in both directions between those two terminals:

"Normal" SCRs, lacking this internal resistor, are sometimes referred to as sensitive gate
SCRs due to their ability to be triggered by the slightest positive gate signal.

The test circuit for an SCR is both practical as a diagnostic tool for checking suspected
SCRs and also an excellent aid to understanding basic SCR operation. A DC voltage source
is used for powering the circuit, and two pushbutton switches are used to latch and unlatch
the SCR, respectively:
Actuating the normally-open "on" pushbutton switch connects the gate to the anode,
allowing current from the negative terminal of the battery, through the cathode-gate PN
junction, through the switch, through the load resistor, and back to the battery. This gate
current should force the SCR to latch on, allowing current to go directly from cathode to
anode without further triggering through the gate. When the "on" pushbutton is released,
the load should remain energized.

Pushing the normally-closed "off" pushbutton switch breaks the circuit, forcing current
through the SCR to halt, thus forcing it to turn off (low-current dropout).

If the SCR fails to latch, the problem may be with the load and not the SCR. There is a
certain minimum amount of load current required to hold the SCR latched in the "on" state.
This minimum current level is called the holding current. A load with too great a resistance
value may not draw enough current to keep an SCR latched when gate current ceases, thus
giving the false impression of a bad (unlatchable) SCR in the test circuit. Holding current
values for different SCRs should be available from the manufacturers. Typical holding
current values range from 1 milliamp to 50 milliamps or more for larger units.

For the test to be fully comprehensive, more than the triggering action needs to be tested.
The forward breakover voltage limit of the SCR could be tested by increasing the DC
voltage supply (with no pushbuttons actuated) until the SCR latches all on its own. Beware
that a breakover test may require very high voltage: many power SCRs have breakover
voltage ratings of 600 volts or more! Also, if a pulse voltage generator is available, the
critical rate of voltage rise for the SCR could be tested in the same way: subject it to
pulsing supply voltages of different V/time rates with no pushbutton switches actuated and
see when it latches.

In this simple form, the SCR test circuit could suffice as a start/stop control circuit for a DC
motor, lamp, or other practical load:
Another practical use for the SCR in a DC circuit is as a crowbar device for overvoltage
protection. A "crowbar" circuit consists of an SCR placed in parallel with the output of a
DC power supply, for the purpose of placing a direct short-circuit on the output of that
supply to prevent excessive voltage from reaching the load. Damage to the SCR and power
supply is prevented by the judicious placement of a fuse or substantial series resistance
ahead of the SCR to limit short-circuit current:

Some device or circuit sensing the output voltage will be connected to the gate of the SCR,
so that when an overvoltage condition occurs, voltage will be applied between the gate and
cathode, triggering the SCR and forcing the fuse to blow. The effect will be approximately
the same as dropping a solid steel crowbar directly across the output terminals of the power
supply, hence the name of the circuit.

Most applications of the SCR are for AC power control, despite the fact that SCRs are
inherently DC (unidirectional) devices. If bidirectional circuit current is required, multiple
SCRs may be used, with one or more facing each direction to handle current through both
half-cycles of the AC wave. The primary reason SCRs are used at all for AC power control
applications is the unique response of a thyristor to an alternating current. As we saw in the
case of the thyratron tube (the electron tube version of the SCR) and the DIAC, a hysteretic
device triggered on during a portion of an AC half-cycle will latch and remain on
throughout the remainder of the half-cycle until the AC current decreases to zero, as it must
to begin the next half-cycle. Just prior to the zero-crossover point of the current waveform,
the thyristor will turn off due to insufficient current (this behavior is also known as natural
commutation) and must be fired again during the next cycle. The result is a circuit current
equivalent to a "chopped up" sine wave. For review, here is the graph of a DIAC's response
to an AC voltage whose peak exceeds the breakover voltage of the DIAC:

With the DIAC, that breakover voltage limit was a fixed quantity. With the SCR, we have
control over exactly when the device becomes latched by triggering the gate at any point in
time along the waveform. By connecting a suitable control circuit to the gate of an SCR, we
can "chop" the sine wave at any point to allow for time-proportioned power control to a
load.

Take the following circuit as an example. Here, an SCR is positioned in a circuit to control
power to a load from an AC source:

Being a unidirectional (one-way) device, at most we can only deliver half-wave power to
the load, in the half-cycle of AC where the supply voltage polarity is positive on the top and
negative on the bottom. However, for demonstrating the basic concept of time-proportional
control, this simple circuit is better than one controlling full-wave power (which would
require two SCRs).
With no triggering to the gate, and the AC source voltage well below the SCR's breakover
voltage rating, the SCR will never turn on. Connecting the SCR gate to the anode through a
normal rectifying diode (to prevent reverse current through the gate in the event of the SCR
containing a built-in gate-cathode resistor), will allow the SCR to be triggered almost
immediately at the beginning of every positive half-cycle:

We can delay the triggering of the SCR, however, by inserting some resistance into the gate
circuit, thus increasing the amount of voltage drop required before there is enough gate
current to trigger the SCR. In other words, if we make it harder for electrons to flow
through the gate by adding a resistance, the AC voltage will have to reach a higher point in
its cycle before there will be enough gate current to turn the SCR on. The result looks like
this:
With the half-sine wave chopped up to a greater degree by delayed triggering of the SCR,
the load receives less average power (power is delivered for less time throughout a cycle).
By making the series gate resistor variable, we can make adjustments to the time-
proportioned power:

Unfortunately, this control scheme has a significant limitation. In using the AC source
waveform for our SCR triggering signal, we limit control to the first half of the waveform's
half-cycle. In other words, there is no way for us to wait until after the wave's peak to
trigger the SCR. This means we can turn down the power only to the point where the SCR
turns on at the very peak of the wave:

Raising the trigger threshold any more will cause the circuit to not trigger at all, since not
even the peak of the AC power voltage will be enough to trigger the SCR. The result will
be no power to the load.

An ingenious solution to this control dilemma is found in the addition of a phase-shifting


capacitor to the circuit:

The smaller waveform shown on the graph is voltage across the capacitor. For the sake of
illustrating the phase shift, I'm assuming a condition of maximum control resistance where
the SCR is not triggering at all and there is no load current, save for what little current goes
through the control resistor and capacitor. This capacitor voltage will be phase-shifted
anywhere from 0o to 90o lagging behind the power source AC waveform. When this phase-
shifted voltage reaches a high enough level, the SCR will trigger.

Assuming there is periodically enough voltage across the capacitor to trigger the SCR, the
resulting load current waveform will look something like this:

Because the capacitor waveform is still rising after the main AC power waveform has
reached its peak, it becomes possible to trigger the SCR at a threshold level beyond that
peak, thus chopping the load current wave further than it was possible with the simpler
circuit. In reality, the capacitor voltage waveform is a bit more complex that what is shown
here, its sinusoidal shape distorted every time the SCR latches on. However, what I'm
trying to illustrate here is the delayed triggering action gained with the phase-shifting RC
network, and so a simplified, undistorted waveform serves the purpose well.

SCRs may also be triggered, or "fired," by more complex circuits. While the circuit
previously shown is sufficient for a simple application like a lamp control, large industrial
motor controls often rely on more sophisticated triggering methods. Sometimes, pulse
transformers are used to couple a triggering circuit to the gate and cathode of an SCR to
provide electrical isolation between the triggering and power circuits:
When multiple SCRs are used to control power, their cathodes are often not electrically
common, making it difficult to connect a single triggering circuit to all SCRs equally. An
example of this is the controlled bridge rectifier shown here:

In any bridge rectifier circuit, the rectifying diodes (or in this case, the rectifying SCRs)
must conduct in opposite pairs. SCR1 and SCR3 must be fired simultaneously, and likewise
SCR2 and SCR4 must be fired together as a pair. As you will notice, though, these pairs of
SCRs do not share the same cathode connections, meaning that it would not work to simply
parallel their respective gate connections and connect a single voltage source to trigger
both:

Although the triggering voltage source shown will trigger SCR4, it will not trigger SCR2
properly because the two thyristors do not share a common cathode connection to reference
that triggering voltage. Pulse transformers connecting the two thyristor gates to a common
triggering voltage source will work, however:

Bear in mind that this circuit only shows the gate connections for two out of the four SCRs.
Pulse transformers and triggering sources for SCR 1 and SCR3, as well as the details of the
pulse sources themselves, have been omitted for the sake of simplicity.

Controlled bridge rectifiers are not limited to single-phase designs. In most industrial
control systems, AC power is available in three-phase form for maximum efficiency, and
solid-state control circuits are built to take advantage of that. A three-phase controlled
rectifier circuit built with SCRs, without pulse transformers or triggering circuitry shown,
would look like this:

 REVIEW:
 A Silicon-Controlled Rectifier, or SCR, is essentially a Shockley diode with an extra
terminal added. This extra terminal is called the gate, and it is used to trigger the
device into conduction (latch it) by the application of a small voltage.
 To trigger, or fire, an SCR, voltage must be applied between the gate and cathode,
positive to the gate and negative to the cathode. When testing an SCR, a momentary
connection between the gate and anode is sufficient in polarity, intensity, and
duration to trigger it.
 SCRs may be fired by intentional triggering of the gate terminal, excessive voltage
(breakdown) between anode and cathode, or excessive rate of voltage rise between
anode and cathode. SCRs may be turned off by anode current falling below the
holding current value (low-current dropout), or by "reverse-firing" the gate
(applying a negative voltage to the gate). Reverse-firing is only sometimes
effective, and always involves high gate current.
 A variant of the SCR, called a Gate-Turn-Off thyristor (GTO), is specifically
designed to be turned off by means of reverse triggering. Even then, reverse
triggering requires fairly high current: typically 20% of the anode current.
 SCR terminals may be identified by a continuity meter: the only two terminals
showing any continuity between them at all should be the gate and cathode. Gate
and cathode terminals connect to a PN junction inside the SCR, so a continuity
meter should obtain a diode-like reading between these two terminals with the red
(+) lead on the gate and the black (-) lead on the cathode. Beware, though, that some
large SCRs have an internal resistor connected between gate and cathode, which
will affect any continuity readings taken by a meter.
 SCRs are true rectifiers: they only allow current through them in one direction. This
means they cannot be used alone for full-wave AC power control.
 If the diodes in a rectifier circuit are replaced by SCRs, you have the makings of a
controlled rectifier circuit, whereby DC power to a load may be time-proportioned
by triggering the SCRs at different points along the AC power waveform.

The TRIAC

SCRs are unidirectional (one-way) current devices, making them useful for controlling DC
only. If two SCRs are joined in back-to-back parallel fashion just like two Shockley diodes
were joined together to form a DIAC, we have a new device known as the TRIAC:
Because individual SCRs are more flexible to use in advanced control systems, they are
more commonly seen in circuits like motor drives, while TRIACs are usually seen in
simple, low-power applications like household dimmer switches. A simple lamp dimmer
circuit is shown here, complete with the phase-shifting resistor-capacitor network necessary
for after-peak firing.

TRIACs are notorious for not firing symmetrically. This means they usually won't trigger at
the exact same gate voltage level for one polarity as for the other. Generally speaking, this
is undesirable, because unsymmetrical firing results in a current waveform with a greater
variety of harmonic frequencies. Waveforms that are symmetrical above and below their
average centerlines are comprised of only odd-numbered harmonics. Unsymmetrical
waveforms, on the other hand, contain even-numbered harmonics (which may or may not
be accompanied by odd-numbered harmonics as well).

In the interest of reducing total harmonic content in power systems, the fewer and less
diverse the harmonics, the better -- one more reason why individual SCRs are favored over
TRIACs for complex, high-power control circuits. One way to make the TRIAC's current
waveform more symmetrical is to use a device external to the TRIAC to time the triggering
pulse. A DIAC placed in series with the gate does a fair job of this:

DIAC breakover voltages tend to be much more symmetrical (the same in one polarity as
the other) than TRIAC triggering voltage thresholds. Since the DIAC prevents any gate
current until the triggering voltage has reached a certain, repeatable level in either direction,
the firing point of the TRIAC from one half-cycle to the next tends to be more consistent,
and the waveform more symmetrical above and below its centerline.

Practically all the characteristics and ratings of SCRs apply equally to TRIACs, except that
TRIACs of course are bidirectional (can handle current in both directions). Not much more
needs to be said about this device except for an important caveat concerning its terminal
designations.

From the equivalent circuit diagram shown earlier, one might think that main terminals 1
and 2 were interchangeable. They are not! Although it is helpful to imagine the TRIAC as
being composed of two SCRs joined together, it in fact is constructed from a single piece of
semiconducting material, appropriately doped and layered. The actual operating
characteristics may differ slightly from that of the equivalent model.

This is made most evident by contrasting two simple circuit designs, one that works and
one that doesn't. The following two circuits are a variation of the lamp dimmer circuit
shown earlier, the phase-shifting capacitor and DIAC removed for simplicity's sake.
Although the resulting circuit lacks the fine control ability of the more complex version
(with capacitor and DIAC), it does function:

Suppose we were to swap the two main terminals of the TRIAC around. According to the
equivalent circuit diagram shown earlier in this section, the swap should make no
difference. The circuit ought to work:

However, if this circuit is built, it will be found that it does not work! The load will receive
no power, the TRIAC refusing to fire at all, no matter how low or high a resistance value
the control resistor is set to. The key to successfully triggering a TRIAC is to make sure the
gate receives its triggering current from the main terminal 2 side of the circuit (the main
terminal on the opposite side of the TRIAC symbol from the gate terminal). Identification
of the MT1 and MT2 terminals must be done via the TRIAC's part number with reference to
a data sheet or book.
 REVIEW:
 A TRIAC acts much like two SCRs connected back-to-back for bidirectional (AC)
operation.
 TRIAC controls are more often seen in simple, low-power circuits than complex,
high-power circuits. In large power control circuits, multiple SCRs tend to be
favored.
 When used to control AC power to a load, TRIACs are often accompanied by
DIACs connected in series with their gate terminals. The DIAC helps the TRIAC
fire more symmetrically (more consistently from one polarity to another).
 Main terminals 1 and 2 on a TRIAC are not interchangeable.
 To successfully trigger a TRIAC, gate current must come from the main terminal 2
(MT2) side of the circuit!
Optothyristors

Like bipolar transistors, SCRs and TRIACs are also manufactured as light-sensitive
devices, the action of impinging light replacing the function of triggering voltage.

Optically-controlled SCRs are often known by the acronym LASCR, or Light Activated
SCR. Its symbol, not surprisingly, looks like this:

Optically-controlled TRIACs don't receive the honor of having their own acronym, but
instead are humbly known as opto-TRIACs. Their schematic symbol looks like this:

Optothyristors (a general term for either the LASCR or the opto-TRIAC) are commonly
found inside sealed "optoisolator" modules.

The Unijunction Transistor (UJT) -- PENDING

Programmable Unijunction Transistors (PUTs).

 REVIEW:


The Silicon-Controlled Switch (SCS)

If we take the equivalent circuit for an SCR and add another external terminal, connected to
the base of the top transistor and the collector of the bottom transistor, we have a device
known as a silicon-controlled-switch, or SCS:
This extra terminal allows more control to be exerted over the device, particularly in the
mode of forced commutation, where an external signal forces it to turn off while the main
current through the device has not yet fallen below the holding current value. Consider the
following circuit:

When the "on" pushbutton switch is actuated, there is a voltage applied between the
cathode gate and the cathode, forward-biasing the lower transistor's base-emitter junction,
and turning it on. The top transistor of the SCS is ready to conduct, having been supplied
with a current path from its emitter terminal (the SCS's anode terminal) through resistor R 2
to the positive side of the power supply. As in the case of the SCR, both transistors turn on
and maintain each other in the "on" mode. When the lower transistor turns on, it conducts
the motor's load current, and the motor starts and runs.
The motor may be stopped by interrupting the power supply, as with an SCR, and this is
called natural commutation. However, the SCS provides us with another means of turning
off: forced commutation by shorting the anode terminal to the cathode. If this is done (by
actuating the "off" pushbutton switch), the upper transistor within the SCS will lose its
emitter current, thus halting current through the base of the lower transistor. When the
lower transistor turns off, it breaks the circuit for base current through the top transistor
(securing its "off" state), and the motor (making it stop). The SCS will remain in the off
condition until such time that the "on" pushbutton switch is re-actuated.

 REVIEW:
 A silicon-controlled switch, or SCS, is essentially an SCR with an extra gate
terminal.
 Typically, the load current through an SCS is carried by the anode gate and cathode
terminals, with the cathode gate and anode terminals sufficing as control leads.
 An SCS is turned on by applying a positive voltage between the cathode gate and
cathode terminals. It may be turned off (forced commutation) by applying a
negative voltage between the anode and cathode terminals, or simply by shorting
those two terminals together. The anode terminal must be kept positive with respect
to the cathode in order for the SCS to latch.
Field-effect-controlled thyristors

Two relatively recent technologies designed to reduce the "driving" (gate trigger current)
requirements of classic thyristor devices are the MOS-gated thyristor and the MOS
Controlled Thyristor, or MCT.

The MOS-gated thyristor uses a MOSFET to initiate conduction through the upper (PNP)
transistor of a normal thyristor structure, thus triggering the device. Since a MOSFET
requires negligible current to "drive" (cause it to saturate), this makes the thyristor as a
whole very easy to trigger:

Given the fact that ordinary SCRs are quite easy to "drive" as it is, the practical advantage
of using an even more sensitive device (a MOSFET) to initiate triggering is debatable.
Also, placing a MOSFET at the gate input of the thyristor now makes it impossible to turn
it off by a reverse-triggering signal. Only low-current dropout can make this device stop
conducting after it has been latched.

A device of arguably greater value would be a fully-controllable thyristor, whereby a small


gate signal could both trigger the thyristor and force it to turn off. Such a device does exist,
and it is called the MOS Controlled Thyristor, or MCT. It uses a pair of MOSFETs
connected to a common gate terminal, one to trigger the thyristor and the other to
"untrigger" it:
A positive gate voltage (with respect to the cathode) turns on the upper (N-channel)
MOSFET, allowing base current through the upper (PNP) transistor, which latches the
transistor pair in an "on" state. Once both transistors are fully latched, there will be little
voltage dropped between anode and cathode, and the thyristor will remain latched so long
as the controlled current exceeds the minimum (holding) current value. However, if a
negative gate voltage is applied (with respect to the anode, which is at nearly the same
voltage as the cathode in the latched state), the lower MOSFET will turn on and "short"
between the lower (NPN) transistor's base and emitter terminals, thus forcing it into cutoff.
Once the NPN transistor cuts off, the PNP transistor will drop out of conduction, and the
whole thyristor turns off. Gate voltage has full control over conduction through the MCT:
to turn it on and to turn it off.

This device is still a thyristor, though. If there is zero voltage applied between gate and
cathode, neither MOSFET will turn on. Consequently, the bipolar transistor pair will remain
in whatever state it was last in (hysteresis). So, a brief positive pulse to the gate turns the
MCT on, a brief negative pulse forces it off, and no applied gate voltage lets it remain in
whatever state it is already in. In essence, the MCT is a latching version of the IGBT
(Insulated Gate Bipolar Transistor).

 REVIEW:
 A MOS-gated thyristor uses an N-channel MOSFET to trigger a thyristor, resulting
in an extremely low gate current requirement.

A MOS Controlled Thyristor, or MCT, uses two MOSFETS to exert full control over the
thyristor. A positive gate voltage triggers the device, while a negative gate voltage forces it
to turn off. Zero gate voltage allows the thyristor to remain in whatever state it was
previously in (off, or latched on).

You might also like