Difference between revisions of "Electronics primer"

From Course Wiki
Jump to: navigation, search
(Solving problems: strategy)
(Thevenin equivalent circuits)
 
(242 intermediate revisions by 6 users not shown)
Line 1: Line 1:
 +
[[Category:20.309]]
 +
[[Category:Electronics]]
 
{{Template:20.309}}
 
{{Template:20.309}}
 +
[[Image:xkcd_circuit_diagram.png|thumb|275 px|right|Circuit diagram according to [http://xkcd.com/730/ xkcd].]]
  
Reference format wiki markup: <ref name="Invitrogen Spetral Viewer">[http://www.invitrogen.com/site/us/en/home/support/Research-Tools/Fluorescence-SpectraViewer.html Invitrogen Spectral Viewer]</ref>
+
==Overview==
  
==Overview: Should I read this thing?==
+
[[Image:309_epd_resistor-fully-labeled.png|thumb|right|Resistor symbol.]]
 +
The physical quantities that characterize many types of systems are defined at all points in space and time. In the case of electric circuits, the quantities of interest are charge, electric field, and magnetic field. Maxwell's equations describe the relationship between the three quantities as a function of space and time. Using Maxwell's equations to compute the field and charge at every point is a very thorough approach to analyzing electrical systems. However, such a detailed model includes far too much information to be useful for understanding or designing most circuits made from standard electronic components like resistors and capacitors. Analysis can be dramatically simplified by using a ''lumped element'' circuit model. Lumped element models characterize a system only at a small number of well-defined points by collapsing regions of space into ideal lumped elements. Regions of space that do not significantly impact system behavior are ignored. 
  
This primer is intended to quickly get electronics newbies comfortable with circuit interpretation by providing clear definitions and systematic mathematical approaches, without unnecessary detail.
+
Lumped electronic elements are idealized abstractions of physical components like resistors and capacitors. A lumped element circuit model represents a system as a network of elements connected together by wires. Each element has two ''state variables'': current and voltage. (The lumped element model for electronic components assumes that magnetic interaction between elements are negligible.) It is possible to create lumped element models of many other types of systems described by two state variables, including thermal, acoustic, hydraulic, and mechanical systems.  
  
If you don’t feel comfortable defining an open or short circuit across arbitrary nodes, aren’t sure how to keep your voltage and current signs consistent, and perhaps even barely remember what series versus parallel means, then this guide is for you. You probably want to read all of it, preferably sections 1-4 before the second lecture on electronics (and section 5 in concert with the first electronics problem set). You’ll get more out of lecture if you’re not struggling with the basics.
+
Elements are bounded by terminals. In elements with two terminals, one is typically designated plus (+) and the other minus (-). Every two-terminal element has a single voltage across it and a single current flow through it. The relationship between current through the element and the voltage difference across the element's terminals is defined by a constitutive equation<ref name="CircuitAssumptions"> For more about circuit assumptions and abstraction layers, see Agarwal 1.1-1.4: ''The power of abstraction'', ''The lumped circuit abstraction'', ''The lumped matter discipline'', and ''Limitations of the lumped circuit abstraction''.</ref>.
  
If you do feel comfortable with the basic concepts and math in theoretical form, but have no clue how these relate to instruments, breadboards, and what leads are touching where, then you can probably skip this guide and focus your efforts on completing Module 0 instead.  
+
Each type of element has an associated graphic symbol. The symbol for a resistor is shown at right. The constitutive equation for an ideal resistor is given by Ohm's law, ''v''=''iR''.
  
References to the relevant sections of Agarwal and Lang (6.002 textbook, 2005 edition) are included for those seeking to further solidify their understanding.
+
==Review of electrical units and equations==
  
==Motivation: Where is this all leading and why should I care?==
+
===Voltage===
  
Circuit representations are useful for solving problems in many engineering domains -- electrical, mechanical, thermal, hydraulic, etc.  Any system of linear ordinary differential equations is amenable to circuit analysis. For simplicity, we’ll start with circuit elements whose behavior is not frequency dependent, namely resistors and sources, and build up to more complex systems involving capacitors and inductors in lecture/lab and perhaps someday a subsequent primer.
+
[[Image:309_epd_grouds-example.png|thumb|right|250px|Multi-component circuit with ground nodes emphasized.]]
  
==Defining essential circuit elements and variables: v, i, R==
+
Voltage quantifies the electrical potential difference between two terminals. Higher voltages attract electrons more intensely and thus result in larger current flows. Voltage is usually represented by the variable ''v''.
  
===Primary definitions===
+
Electric potential difference has the units of energy divided by charge and is measured in volts, abbreviated with a capital V. Energy is measured in joules (J) and electric charge is measured in coulombs (C)<ref>One mole of electrons has a charge of about 96,500 coulombs.</ref>. 1 V = 1  J/C = 1 kg m<sup>2</sup> / s<sup>2</sup> C.
  
Table 1 lists the name, symbol, constitutive relation and its graphical representation for each of four simple idealized elements: a voltage source, a current source, a resistor, and a conductor. Each constitutive equation and associated plot relate voltage (''v'') across an element and current (''i'') through that element.  
+
Voltage is always measured relative to a reference point. It is useful to designate a particular node as the zero point, called the ''ground node''. You may designate any node at all as ground; however, some choices are more convenient than others. In a circuit with a single voltage or current source, the best choice is usually the negative terminal of the source. The voltage across an element in a network can be found by subtracting the node voltage at its minus terminal from the node voltage at its plus terminal.  
  
Current reflects the flow rate of charge carriers such as electrons and by convention is positive in the direction of positive charge flow, which means the electrons are actually flowing in the negative direction. Voltage reflects the electrical potential difference between two points. The higher the voltage, the greater the current flow (always down the potential gradient, because electrons are attracted up the potential gradient) through a given element. To make circuit analysis tenable, each element is treated as having a uniform ''v'' and ''i''.<ref name="CircuitAssumptions">For more about circuit assumptions and abstraction layers, see Agarwal 1.1-1.4: ''The power of abstraction'', ''The lumped circuit abstraction'', ''The lumped matter discipline'', and ''Limitations of the lumped circuit abstraction''.</ref>
+
Another possible choice of reference points is ''earth ground''. Earth ground is the potential at which we spend most of our days &mdash; the potential of a conductor driven a few feet into the soil. If you are at earth ground potential, the doorknob doesn't shock you when you go to open the door. The ground node in most circuits is usually at a potential close to earth ground, but not always. The third conductor on electrical plugs is connected to earth ground. Since we spend most of our time near earth ground potential, connecting the outer case of electrical equipment to earth ground helps prevent shocks. Earth ground and the circuit ground node may or may not be connected together.
  
[[Image:309_epd_Table1.jpg|thumb|center|600px|'''Table 1: Constitutive relations for ideal elements.''']]
+
===Current===
<br style="clear:both;"/>
+
Current measures the flow rate of charge through an element. The SI unit for current is amperes, abbreviated amps of just A. 1 A = 1 C/s. A positive current flows from the plus terminal to the minus. The variable ''i'' is usually used to represent current.
  
[[Image:309_epd_resistor-fully-labeled.png|thumb|225px|right|'''Figure 1: Resistor with terminal parameters defined.''' The voltage across and current through the element are shown.]]
+
Skip the following paragraph if you want to keep things simple.
An element is bracketed by terminals, one designated (+) and the other (-). Terminals may be shown as filled or open circles to indicate whether or not that terminal connects to another electrical element. (Open circles represent no connection.) Pairs of terminals are sometimes called ports, whether for a single element or across multiple elements. A port approach is particularly useful for abstracting away (black-boxing) multiple circuit elements and treating them as an equivalent element with defined inputs and/or outputs. We'll return to this idea in sections 5.4 and 5.6.  
+
  
A voltage ''v'' and current ''i'' can be defined at any terminal. For terminals A and B in Figure 1, the element voltage <math>v = v_{AB} = v_A - v_B</math> and the element current <math>i = i_A = i_B</math>. Here, <math>v_A</math> and <math>v_B</math> are the absolute potentials (defined relative to a reference), and ''v'' is the potential difference between the two terminals.  
+
The conventional definition of current is the flow rate of positive charge carriers from the plus terminal to the minus terminal. In most of the components you will work with, such as wires, capacitors, and resistors, the charge carriers are negatively charged electrons. Electrons flowing from the minus to the plus terminal create a current with a positive sign. Current in semiconductors is carried by both positive and negative carriers. This is confusing. It's all Benjamin Franklin's fault<ref>http://www.physicsclassroom.com/class/circuits/u9l2c.cfm</ref>. The best approach is not to think about it at all when you are solving circuit problems. Stick to the convention that positive current flows from plus to minus. It doesn't matter for most circuits what the sign of the charge carriers is. If you are doing a gel electrophoresis, on the other hand, the sign of the charge carriers is crucial.
<br style="clear:both;"/> 
+
  
Expanding upon Table 1,  
+
===Resistance===
 +
Resistance has units of ohms, sometimes abbreviated with the greek letter &Omega;. 1 &Omega; = 1 V/A. Resistance is usually denoted by the letter R.
  
#An ideal voltage source maintains a set potential difference across its terminals no matter what current is being drawn through it. Thus, its voltage versus current graph is simply a horizontal line. A real source such as a battery does not exhibit this behavior. However, such non-ideal elements can be modeled as combinations of ideal linear elements, as described in section 5.6.
+
===Power===
#Similarly, an ideal current source provides a set current no matter what potential difference exists across its terminals. Real sources include op-amps.
+
Electric power is equal to voltage multiplied by current, ''P'' = ''v i''. Power has units of energy per time called watts, abbreviated with a capital W. 1 W = 1V x 1A = 1 J / s.
#Resistance reflects the tendency of an element to draw current and also its tendency to dissipate energy. A linear resistor obeys Ohm’s law, <math>v = iR</math>, at any instant in time. Real resistors are well represented by the model resistor equation only so long as the current and voltage do not exceed their maximum combined ratings.
+
#Ideal conductors have no resistance, meaning that no potential difference appears ''across'' them for any value of current flow ''through'' them. No voltage drop in turn results in no energy dissipation. Ideal conductors are a fine model for real wires.
+
  
Elements such as sources and resistors are connected at their terminals by wires in order to form circuits. To make circuit analysis tenable, elements are treated as interacting only through their terminals; in other words, electromagnetic fields associated with a given element are internal to it. When this assumption does not hold, more complex elements called capacitors and inductors can be integrated into the circuit.
+
==Ideal lumped circuit elements==
  
Two limiting cases of the resistor element are useful for solving circuit problems: the open circuit and the short circuit (Fig. 2). Infinite resistance means that no current passes through that element; this is also called an open circuit across said element and is equivalent to an unconnected port. Zero resistance means no potential difference appears across the element; this is also called a short circuit across that element and is equivalent to an ideal wire.
+
[[Image:309_epd_Table1.jpg|325px|thumb|right|Table of ideal elements.]]
 +
The table on the right summarizes the characteristics of four ideal elements: a voltage source, a current source, a resistor, and a conductor. The i-v curve in the fourth column shows the constitutive equation graphically, with voltage plotted on the vertical axis and current on the horizontal axis.  
  
[[Image:309_epd_open-and-short-circuits.png|thumb|550px|center|'''Figure 2: Open and short circuits.''' In the open circuit (left), the maximum potential difference exists across the resistor, and <math>V_{AB} = V</math>. In the short circuit (right), the potential is the same everywhere because a wire connects the two resistor terminals, and <math>V_{AB} = 0</math>.]]
+
An ideal voltage source maintains a constant potential difference across its terminals regardless of the amount of current flowing through it. Thus, its voltage versus current i-v graph is a horizontal line. The current flowing through an ideal current source is constant no matter what potential difference is across its terminals. The plot is a vertical line. Ideal resistors obey Ohm’s law, ''v'' = ''iR'', which is a sloped line on the voltage versus current graph. Ideal conductors have zero resistance. All points on a conductor have the same voltage.  
  
===Illustrative circuit analogies===
+
You can watch a video on ideal circuit elements [https://www.khanacademy.org/science/electrical-engineering/ee-circuit-analysis-topic/circuit-elements/v/ideal-circuit-elements here].
  
We are now in a position to appreciate some analogies to electronic circuits. We’ll only describe in detail the most physically intuitive one, while the rest are listed in Table 2.
+
===Connecting circuit elements===
 +
[[Image:309_epd_node-loop-terminology.png|thumb|250px|right|Circuit illustrating nodes, branches, and loops.]]
  
[[Image:309_epd_Table2.jpg‎|thumb|center|500px|'''Table 2: Circuit variables in other engineering disciplines.''']]
+
Networks of elements are formed by connecting terminals together with wires. A schematic diagram like the one shown at right is one way to specify the connections. Elements interact only through their terminals. Terminals joined together by wires are called nodes. All of the terminals connected to a node have the same voltage, which is unsurprising because they are connected by an an element that by definition has the same potential everywhere. Any path between two independent nodes is called a branch. Any path from one node back around to itself is called a loop.  
  
In a hydraulic system (Fig. 3), pressure differences between two points are analogous to voltages in electronic circuits. A water reservoir is like a voltage source, and water flow is analogous to current flow. The higher the pressure differential (greater amount/height of water in the reservoir), the faster the volumetric flow, all else being equal. Resistance can be thought of as the narrowness of the conduit connecting two points. The narrower the pipe, the slower the flow, all else being equal. Or it can be thought of as a pipe packed with beads. The smaller the beads, or the greater in number, the slower the flow, all else being equal.
+
Terminals may be shown as filled or open circles to indicate whether or not that terminal connects to another element. Open circles represent no connection. Pairs of terminals are sometimes called ''ports''. A port may encompass a single element or multiple elements. Dividing a circuit into groups of elements that interact through their ports can be a useful way to conceptualize and analyze systems.
  
[[Image: 309-epd_reservoir-analogy.png|thumb|center|500px|'''Figure 3: Fluid dynamics analogy for circuits.''' At left, the (pipe) resistance is high, slowing water flow from the reservoir, but the voltage equivalent is also high, quickening water flow compared to a less full reservoir. At right, the parameters are just the opposite.]]
+
==Common circuit patterns==
<br style="clear:both;"/>
+
Two patterns of connected elements come up so frequently that it is worth memorizing them: series and parallel. Many circuit problems can be simplified by recognizing elements within the system that are connected in series or parallel.
  
===Essential units===
+
===Series and parallel connections===
 +
[[Image:309_epd_series-parallel.png|thumb|250px|right|Resistors in series and parallel.]]
  
<math>v = iR</math> <br>  
+
When two elements are joined end to end, they are said to be in series. In this case, the minus terminal of one element is joined to the plus terminal of the second element. The current flowing through each element connected in series connection is identical. The equivalent resistance of two resistors in series is the sum of the resistances: ''R'' = ''R<sub>1</sub>'' + ''R<sub>2''</sub>. Voltage sources connected in series are added.
[Voltage] = [Amps] [Ohms] = A &Omega;<br>
+
Or in a more practical configuration,<br>
+
[Voltage] = [milliAmps] [kiloOhms] = mA k&Omega;
+
  
Another useful quantity that you will encounter is power, or energy transfer per time. Some circuits may be solved by energy balance alone: for example, energy supplied by a voltage source may be dissipated in resistors or stored in capacitors. Note that power out of an element is negative, and power into an element is positive, by convention.<br>
+
In a parallel connection, the plus terminals of two elements are connected together, as are the minus terminals. The voltage across each of the elements in a parallel connection is the same. The equivalent resistance of two resistors in parallel is: ''R'' = <sup>''R<small>1</small>R<small>2</small>''</sup>/<sub>''R<small>1</small>'' + ''R<small>2</small>''</sub>.
<math>p = vi</math> <br>
+
<!-->Voltage sources linked in parallel maintain the potential difference of a single voltage source (so all sources must be identical or things get tricky ), with increased current flow.<ref name="SeriesParallelBatteries">See Agarwal 1.5.1, ''Practical two-terminal elements: Batteries'', for more information.</ref>
[Watts] = [Volts] [Amps] = V A <br>
+
<-->
= [Joule/Coloumb] [Coulomb/second] = [Joule/second] = J/s = W
+
  
==Connecting circuit elements==
+
===Voltage divider===
 +
[[Image:309_epd_basis-voltage-divider.png|150px|thumb|right]]
  
Multiple elements may be joined at their terminals to build a circuit (Fig. 4). The joined terminals are now called nodes. Any path between two (independent -- see below) nodes is called a branch. Any path from one node back around to itself is called a loop. We’ll revisit these terms when we solve circuit problems in section 5.  
+
A very common circuit construct consists of two resistors in series, as shown in the schematic diagram on the right. This type of circuit is called a voltage divider. In a series connection, the same current ''i'' flows through both resistors. The constitutive relation for a resistor is: ''v''=''i''R. Using this equation, the voltage across resistor R<sub>1</sub> is ''i''R<sub>1</sub> and the voltage across R<sub>2</sub> is ''i''R<sub>2</sub>. The minus terminal of R<sub>2</sub> is connected to ground, so the voltage ''V<sub>C</sub>'' at node C is by definition 0 volts. The voltage across R<sub>2</sub> is ''i''R<sub>2</sub>, so the voltage ''V<sub>B</sub>'' at node B is 0+''i''R<sub>2</sub>=''i''R<sub>2</sub>. The ideal source constrains the voltage ''V<sub>A</sub>'' at node A to be V, which must also be equal to the sum of the voltages across the resistors: V = ''i''R<sub>1</sub>+''i''R<sub>2</sub>=''i''(R<sub>1</sub>+R<sub>2</sub>). Substituting for ''V<sub>B</sub>'' into this expression gives an equation for ''V<sub>B</sub>'' in terms of the source voltage V and the resistor values: ''V<sub>B''=V<sup>R<small>2</small></sup>/<sub>R<small>1</small>+R<small>2</small></sub>. If you don't know why the circuit is called a voltage divider, take another look at the equation.
  
[[Image:309_epd_node-loop-terminology.png|thumb|550px|center|'''Figure 4: Circuit illustrating definitions for node, branch, and loop.''']]
+
You can watch a video on voltage dividers [https://www.khanacademy.org/science/electrical-engineering/ee-circuit-analysis-topic/ee-resistor-circuits/v/ee-voltage-divider here].
  
Note that two (or more) apparent nodes that are connected by an ideal conductor are considered equivalent, because the conductor by definition cannot maintain a potential difference. Thus, A and the (A)'s above represent only one independent node, not three, as do B and the (B)'s. The two Steve's are of course independent nodes :)
+
==Solving circuit problems==
  
Now is a good time to emphasize the difference between an element variable and a node variable. The element voltage <math>v_{AB}</math> is the difference between the two node voltages, <math>v_A</math> and <math>v_B</math>. In other words, an element voltage equals the potential difference across two terminals, while a node voltage equals the absolute potential at a terminal. Of course, this "absolute" potential must still be with respect to a common reference point.
+
===Charge and energy conservation===
 +
Charge and energy are conserved in electric circuits. Two equations ensure that both quantities are conserved: Kirchoff’s Current Law (KCL) and Kirchoff’s Voltage Law (KVL).  
  
===Series and parallel===
+
The current law states that the sum of the currents flowing into any node must equal zero. In other words, the sum of currents flowing in must equal the sum of the currents flowing out. Applying KCL to the node in-between two resistors in series demonstrates why the current is equal in both. At the shared node, current flowing out of the first element has nowhere else to go. No other branches exist at that node.
  
When two elements are joined end to end, they are said to be in series. In this case, the (-) terminal of one element is joined to the (+) terminal of the second element. If instead the (+) terminals of the two elements are connected, as are their (-) terminals, then the elements are said to be in parallel.  
+
KVL ensures that energy is conserved. The law states that the sum of the potential difference across elements in a loop is zero. In other words, if you go around a loop, you end up at the same point.
  
[[Image:309_epd_series-parallel.png|thumb|550px|center|'''Figure 5: Resistors in series and parallel (with respect to one voltage source).''']]
+
KVL and KCL also offer another rationale for why the potential must be equal across elements connected in parallel. The two shared nodes form a loop. In order to sum to 0, the potential differences across the two elements must be identical.
  
Resistors in series are additive: <big><math>R = R_1 + R_2</math></big>.
+
===Node-voltage analysis method===
 +
One rigorous way to solve circuit problems is to write the constitutive relation for each element (B equations for B branches), every independent KCL statement (N-1 equations for N nodes), and every independent KVL statement (B-1+N equations for B branches and N nodes). Then algebraic substitutions are made until every element voltage and current is known. Often it is not obvious which substitutions will most quickly yield the quantities that are actually of interest in that problem.
  
Resistors in parallel are inversely additive: <math>\frac{1}{R} = \frac{1}{R_1} + \frac{1}{R_2}</math>, and thus <math>R = \frac{R_1 R_2}{R_1 + R_2}.</math>
+
[[Image:309_epd_voltage-divider-w-currents.png|thumb|right|250px|Voltage divider with current flow convention shown. Current flow in this circuit is clockwise. Hence, <math>i_2</math> is negative.]]
 +
The node-voltage analysis method is a faster universal approach for solving circuit problems. It indirectly incorporates KVL and requires fewer KCL statements. Rarely can there be confusion about which choice of equations will quickly yield meaningful results. (For very complex circuits, linear rather than simple algebraic approaches may be required.)
  
These relationships are consistent with resistance being proportional to length and inversely proportional to area for a given material.  The current that flows through each resistor in series must be identical, while for resistors in parallel, the voltage across each one must be identical. We will learn why when we discuss Kirchoff’s laws in section 5.1.  
+
Recall that while an element voltage is the potential difference across said element, a node voltage is the potential difference between that node and a common reference point – usually ground. (If these definitions don't sound familiar, see Figure 1 under [http://measure.mit.edu/~20.309/wiki/index.php?title=Electronics_Primer#Primary_definitions  Primary definitions] and also review [http://measure.mit.edu/~20.309/wiki/index.php?title=Electronics_Primer#Connecting_circuit_elements  Connecting circuit elements].)
  
Similarly, voltage sources connected in series are additive, with current unaffected. Voltage sources linked in parallel maintain the potential difference of a single voltage source (so all sources must be identical or things get tricky ), with increased current flow.<ref name="SeriesParallelBatteries">See Agarwal 1.5.1, ''Practical two-terminal elements: Batteries'', for more information.</ref>
+
Working with node voltages rather than element voltages makes KVL implicit, as we saw in the voltage divider example above. The basic approach is then to write a single KCL equation per unknown node voltage, while immediately substituting in the element constitutive laws in terms of voltages and resistances.<ref name="FloatingSources"> Slightly more complex approaches must be taken for cases in which a voltage source has a negative terminal that is not connected to ground. See Agarwal 3.3.2, ''The node method: Floating independent voltage source''s.</ref>
  
Series and parallel relationships can be used to reduce complex circuits to more simple ones as an intermediate step in problem-solving.
+
To avoid mistakes when using KCL, it is helpful to adopt a consistent way of writing the node equations. One convenient way to do this is to write an expression of the current flowing ''into'' the node through each connected component, sum the expressions, and set the whole thing equal to zero. The expression for the current flowing through an element into the node is obtained by subtracting the voltage at the node under consideration from the voltage at the far end of the element (i.e. the terminal not connected to the node) and dividing by R. All you have to remember is, "the sum of the currents flowing into the node is zero."
  
===Ground===
+
===Example: measuring voltage===
 +
[[Image:309_epd_voltmeter-voltage-divider.png|thumb|right|250px|Voltage divider with voltmeter attached.]]
 +
Imagine connecting a voltmeter to measure the output of a voltage divider. In order to read the correct potential, the voltmeter must also be connected to ground. A model for the voltmeter includes an ideal meter and a resistance, R<sub>m</sub>. R<sub>m</sub> models the current ''i<sub>3</sub>'' drawn by the voltmeter as it makes its measurement.
  
Note that a voltage, or potential difference, is always a relative description. For example, for a voltage source the nominal (supposed) voltage -- say, 12 V -- is the potential difference between the (+) and (-) terminals. It is usually really convenient when solving circuit problems to define the (-) terminal as zero. (The relative nature of voltage allows this arbitrary choice.) This potential level is called ground. Ground can be considered a reservoir of electrons. By definition, any terminal connected to ground by a wire must also be at 0V. Ground is also the usual reference point for defining terminal/node voltages.
+
Before the meter is connected, KCL states that ''i<sub>1</sub>'' + ''i<sub>2</sub>'' = 0. Assuming R<sub>1</sub>=200&Omega; R<sub>2</sub>=600&Omega;, and V = 12 V, solving for the voltage at node B gives:
  
[[Image:309_epd_grouds-example.png|thumb|center|500px|'''Figure 6: Multi-component circuit with all (dependent) grounded nodes marked.''']]
+
<br />
 +
<math>\frac{12-v_B}{200} + \frac{0-v_B}{600} \to  v_B = 9 \text{V}.</math>
  
However, there is nothing stopping us from defining the (+) and (-) terminals as +6V and -6V, and this may be convenient in certain problems.
+
<br />
 +
After the voltmeter is connected, R<sub>2</sub> and R<sub>m</sub> are in parallel. The KCL equation becomes: ''i<sub>1</sub>'' + ''i<sub>2</sub>'' + ''i<sub>3</sub>'' = 0. If the voltmeter has a resistance of 300&Omega;, the voltage at node B is now:
  
===Understanding equivalent diagram topologies and notations, part 1===
+
<br />
 +
<math>\frac{12-v_B}{200} + \frac{0-v_B}{600} +\frac{0-v_B}{300} \to v_B = 6 V.</math>
  
A common group of elements comprises a voltage source and two resistors in series. For reasons we shall see shortly, this grouping or "primitive" is called a voltage divider. Two common ways to depict this circuit are shown below.
+
<br />
 
+
With R<sub>m</sub>=300&Omega;, the measured value is far from nominal value of 9V before the meter was connected. To make an accurate measurement, the voltmeter must have a resistance much larger than the resistors in the divider. if R<sub>m</sub> is increased an order of magnitude to 3 k&Omega;, <math>v_B</math>will be 8.57 V. Now you can see why voltmeters typically have an equivalent resistance of around 10<sup>9</sup> &Omega;.
[[Image:309_epd_shorthand-voltage-divider.png|thumb|250px|right|'''Figure 7B: Voltage divider circuit, separate grounds.''']]
+
[[Image:309_epd_basis-voltage-divider.png|thumb|250px|left|'''Figure 7A: Voltage divider circuit, complete connections.''']]
+
<br style="clear:both;"/>  
+
 
+
We can see that the voltage at node C is trivially 0, and that at node A is the maximum source voltage, because these are connected to the voltage source by ideal wires. The voltage at node B depends on the relative values of the two resistors.
+
 
+
As an example, let's take ''V'' = 12 V, <math>R_1</math> = 200 &Omega;, and <math>R_2</math> = 600 &Omega;. Each resistor is described by <math>v=iR</math>, where ''i'' must be uniform in this purely series-connected circuit. Ohm's law also applies to the circuit as a whole when using the overall series equivalent resistance <math>R_{eq}</math>, namely the sum of the two resistances. Thus, <math>v = 12 \text{V} = i R_{eq} = i (R_1 + R_2) = i (200 + 600)\Omega</math>, which can be solved for ''i'' = 15 mA.
+
 
+
The first resistor is described by <math>v_{AB}= i R_1 =15 mA * 200 \Omega</math>. Thus, it has an element voltage of <math>v_{AB} = 3 \text{V}</math>, and the node joining the two resistors has node voltage <math>v_B = 12 - 3 = 9 \text{V}</math>. Indeed, solving the second resistor equation <math>v_{BC}= i R_2 =15 \text{mA} * 600 \Omega</math> yields 9 V (with <math>v_C</math>= 0 , of course).
+
 
+
Clearly, a voltage divider "divides" the voltage drop across resistors in series in proportion to their resistances. In sections 5.2-5.3 we will see a more universally applicable way to solve this problem, when we cannot rely on our intuition as much. Then we will imagine connecting measurement devices to the circuit and also further discuss topology.
+
 
+
Some other examples of topogically equivalent circuits are shown below (Fig 8).
+
 
+
[[Image:309_epd_topological-equivalence-3R.png|thumb|350px|right|'''Figure 8B: Topological equivalence, 3 resistors.''']]
+
[[Image:309_epd_topological-equivalence-2R.png|thumb|350px|left|'''Figure 8A: Topological equivalence, 2 resistors.''']]
+
<br style="clear:both;"/>
+
 
+
==Putting it all together: solving circuit problems==
+
 
+
===Solving problems: fundamentals/theory===
+
 
+
Two key relationships provide the foundation for circuit analysis: Kirchoff’s circuit law and Kirchoff’s voltage law, hereafter KCL and KVL.
+
KCL is based on charge conservation: at any node, the sum of the current magnitudes flowing in must equal the sum of the current magnitudes flowing out. When solving problems in which current directions are trivially known, this form of the law works well.
+
 
+
And now we see why current must be equal for elements connected in series: at the shared node, current coming out from the first element must have the same magnitude as current coming into the second element. No other branches exist at that node.
+
 
+
Returning to the issue of signs, current always flows from the (+) to the (-) terminal for an element, that is, from higher to lower voltage (except in a voltage source). However, the terminal definitions may not always be obvious/known in advance of solving the circuit problem. To keep signs straight, it will be convenient to have a systematic relationship between ''i'' and ''v'' signs. This relationship is described in the problem-solving section just below.
+
 
+
KVL is based on conservation of energy: the sum of the element voltages (potential difference across each element) around a loop is zero. Equivalently, the potential difference between the starting and ending node in a loop (that is, the same node) must be zero.
+
 
+
Take, for example, the three node voltage divider from section 4.3. Applying KVL directly in a clockwise direction, <math> v_{AB} + v_{BC} +v_{CA} = 3+ 9 -12 =0.</math> KVL is even more straightforwardly applied using node voltages. Again going clockwise, clearly node voltages <math>v_A - v_B + v_B - v_C - v_A + v_C = 0</math>, without even plugging in numbers.
+
 
+
Now we also see why voltage must be equal for elements connected in parallel: the two shared nodes form a loop, and hence the potential differences across the two elements must be identical.
+
 
+
===Solving problems: strategy===
+
 
+
One rigorous way to solve circuit problems is to write the constitutive relation for each element (B equations for B branches), every independent KCL statement (N-1 equations for N nodes), and every independent KVL statement (B-1+N equations for B branches). Then algbraic substitutions are made until ''every'' element voltage and current is known. Often it is not obvious which substitutions will most quickly yield the quantities that are actually of interest in that problem.
+
 
+
A faster universal approach for solving circuit problems indirectly incorporates KVL and requires fewer KCL statements. Rarely can there be confusion about which choice of equations will quickly yield meaningful results. (For very complex circuits, however, linear rather than simple algebraic approaches may be required.)
+
 
+
Recall that while an element voltage is the potential difference across said element, a node voltage is the potential difference between that node and a common reference point – usually ground. (If these definitions doesn't sound familiar, see Figure 1 under [http://measure.mit.edu/~20.309/wiki/index.php?title=Electronics_Primer#Primary_definitions  Primary definitions] and also review [http://measure.mit.edu/~20.309/wiki/index.php?title=Electronics_Primer#Connecting_circuit_elements  Connecting circuit elements].)
+
 
+
Working with node voltages rather than element voltages makes KVL implicit, as we saw in the voltage divider example above. The basic approach is then to write a single KCL equation per unknown node voltage, while immediately substituting in the element constitutive laws in terms of voltages and resistances.
+
 
+
Now we make our alternative statement of KCL: the sum of all currents flowing into a node is zero, where current now includes both magnitude and sign. For <math>i=v/R</math> relationships, always subtract the close node from the far node &mdash; <math>V_{far} - V_{close}</math> &mdash; in the numerator, and the sign of the calculated current should always come out right. It will be positive if flowing in as assumed, and negative if actually flowing out.* Using this sign convention, a current pointing toward the node is positive, and one pointing away fom the node is negative. However, a voltage source whose far node is the negative terminal must yield a positive current, because a source does work to push charge against the usual direction. [Last part unclear I think. Need to think about it further.]
+
 
+
<font color=red>[Sketch.] Simple voltage divider with one unknown node, currents pointing in, signs of true current labeled.</font color>
+
 
+
<nowiki>*</nowiki> You could alternatively subtract the far node from the close node; the key is to be consistent. May as well memorize what Steve does though!
+
<nowiki>*</nowiki> See Agarwal 3.3.2 for cases where a voltage source has a negative terminal that is not connected to ground. <font color=red>[CHECK THAT FIRST PSET PROBLEM; DON’T REMEMBER HAVING TO THINK ABOUT THE FACT THAT IT WAS FLOATING, JUST SUBTRACT TWO NODES.]</font color>
+
 
+
===Understanding equivalent diagram topologies and notations, part 2===
+
 
+
Let’s return to the two-resistor voltage divider described in section 4.3 above to practice our systematic approach for solving circuits, calling the single non-trivial node of interest B for consistency. We can apply KCL to calculate it:
+
 
+
<math>i_1 + i_2 = 0 = \frac{12-v_B}{200} + \frac{0-v_B}{600} \to \text{nominal}\ v_B = 9 \text{V}.</math>
+
 
+
Now let’s imagine connecting a voltmeter to measure the voltage at node B. To read the correct relative potential, the voltmeter must also be connected at ground. In other words, the second resistor and the voltmeter with associated resistance <math>R_m</math> are in parallel. Before attaching the voltmeter, the access port of interest can be shown as:
+
 
+
[[Image:309_epd_open-port-divider.png|thumb|center|400px|'''Voltage divider with access port for voltmeter highlighted.''']]
+
 
+
After attaching the voltmeter, three equivalent circuit diagrams depicting the system are:
+
 
+
[[Image:309_epd_voltmeter-voltage-divider.png|thumb|center|600px|'''Voltage divider with voltmeter attached.''']]
+
<font color=red>Add i1-i3 current flows to first diagram in the triptych</font color>
+
 
+
The voltmeter must have a high enough resistance that it does not draw current and substantially affect <math>v_B</math>. If instead it has a comparable resistance, say 300 &Omega;, let’s calculate the effect on <math>v_B</math> using our usual KCL approach:
+
 
+
<math>i_1 + i_2 + i_3 = 0 = \frac{12-v_B}{200} + \frac{0-v_B}{600} +\frac{0-v_B}{300} \to \text{nominal}\ v_B = 6 V.</math>
+
 
+
This value is unacceptably far off from our nominal 9 V. However, if <math>R_m</math> is increased an order of magnitude, to 3 k&Omega;, <math>v_B</math>is a respectable 8.57 V. (Try it!)
+
 
+
Wouldn't it be nice to calculate <math>v_B</math> for a given <math>R_m</math> without repeating the full calculation each time? The next section describes just such a method.
+
  
 
===Thevenin equivalent circuits===
 
===Thevenin equivalent circuits===
Line 187: Line 122:
 
''Topological'' equivalence between circuits implies that wires have been moved around or added in place of shorthand. There is also such a thing as ''functional'' equivalence between circuits. In fact, any complex circuit consisting only of voltage sources and resistors can be reduced to a circuit with only a single voltage source and resistor! This equivalent circuit cannot be empirically distinguished from the multi-component one.
 
''Topological'' equivalence between circuits implies that wires have been moved around or added in place of shorthand. There is also such a thing as ''functional'' equivalence between circuits. In fact, any complex circuit consisting only of voltage sources and resistors can be reduced to a circuit with only a single voltage source and resistor! This equivalent circuit cannot be empirically distinguished from the multi-component one.
  
[[Image:309_epd_short-circuit-voltage-divider.png|thumb|right|300px|'''Short circuit current for voltage divider.''']]
 
 
Let’s return to the voltage divider example with a port across the second resistor. We want to know how the circuit as a whole responds at this access port when there is no load on it, i.e., no element in which energy can be dissipated. We calculate the equivalent circuit by three steps.
 
Let’s return to the voltage divider example with a port across the second resistor. We want to know how the circuit as a whole responds at this access port when there is no load on it, i.e., no element in which energy can be dissipated. We calculate the equivalent circuit by three steps.
  
First we calculate the open circuit voltage (<math>V_{oc}</math>) at that port, namely 9 V as above in section 5.3.  
+
First we calculate the open circuit voltage (<math>V_{oc}</math>) at that port, namely 9 V as [http://measure.mit.edu/~20.309/wiki/index.php?title=Electronics_Primer#Example:_measuring_voltage  above].  
  
Next we calculate the short circuit current. In this hypothetical, the attached voltmeter has no internal resistance and behaves as a wire (see image). In turn, <math>v_B</math> is shorted to ground, and current magnitude is dependent only on <math>R_1</math>, not both <math>R_1</math> and <math>R_2</math>. Hence, the short circuit current <math>i_{sc} = 12 \text{V}/200 \Omega = 0.06 \text{A} = 60 \text{mA} </math>.
+
Next we calculate the short circuit current. In this hypothetical situation, the attached voltmeter has no internal resistance and behaves as a wire (Fig 12). In turn, <math>v_B</math> is shorted to ground, and current magnitude is dependent only on <math>R_1</math>, not both <math>R_1</math> and <math>R_2</math>. Hence, the short circuit current <math>I_{sc} = 12 \text{V}/200 \Omega = 0.06 \text{A} = 60 \text{mA} </math>.  
\OmegaOhms.</math>
+
  
The Thevenin equivalent circuit is thus:
+
Finally, the Thevenin resistance <math>R_{Th} = \frac{V_{oc}}{I_{sc}} = \frac{9 \text{V}}{0.06 \text{A}} = 150 \Omega</math>. The Thevenin equivalent circuit is shown in Figure 13.
 
+
<font color=red>[Sketch 9V source and 150 Ohm resistor.]</font color>
+
  
 
Now how does this two-component circuit reduction help us? The voltmeter (or any element!) can be connected in series with <math>R_{Th}</math>, and <math>v_B</math> calculated immediately by the simple voltage divider relation, <math>\frac{V R_m}{R_{tot}}.</math>
 
Now how does this two-component circuit reduction help us? The voltmeter (or any element!) can be connected in series with <math>R_{Th}</math>, and <math>v_B</math> calculated immediately by the simple voltage divider relation, <math>\frac{V R_m}{R_{tot}}.</math>
  
An alternative and direct calculation of <math>R_{Th}</math> is the total resistance at the port as viewed looking into the <font color=red>[?not sure how to define direction here]</font color> circuit: here, R1 and R2 in parallel.
+
An alternative and direct calculation of <math>R_{Th}</math> is the total resistance at the port as viewed looking into the circuit (with voltage sources changed to wires): here, <math>R_1</math> and <math>R_2</math> in parallel.
 +
 
 +
[[Image:309_epd_short-circuit-voltage-divider.png|thumb|left|300px|'''Figure 12:''' Short circuit current for voltage divider.]]
 +
[[Image:309_epd_Thevenin-equivalent.png|thumb|center|400px|'''Figure 13: Thevenin reduction of voltage divider.''' The Thevenin equivalent of the previously described voltage divider is shown (left) with an open port ready to be connected to a voltmeter with arbitrary resistance (right).]]
 +
 
 
<br style="clear:both;"/>
 
<br style="clear:both;"/>
  
 
===Alternative problem-solving strategies ===
 
===Alternative problem-solving strategies ===
  
Circuit problems can be approached in many ways<ref name="AlternativeProblemApproaches">Overall, see chapters 2 (''Resistive Networks'') and 3 (''Network Theorems'') of Agarwal. Section 2.4 (''Intuitive method of circuit analysis: series and parallel simplification'') treats series/parallel reductions, section 3.5 (''Superposition'') treats superposition, etc.</ref> and in the interests of time we will primarily treat the approach that (nearly) universally works, namely the node voltage/KCL method described above. Let’s briefly look at one example of another approach.
+
Circuit problems can be approached in many ways<ref name="AlternativeProblemApproaches">Overall, see chapters 2 (''Resistive Networks'') and 3 (''Network Theorems'') of Agarwal. Section 2.4 (''Intuitive method of circuit analysis: series and parallel simplification'') treats series/parallel reductions, section 3.5 (''Superposition'') treats superposition, etc.</ref> and in the interest of time we will primarily treat the approach that (nearly) universally works, namely the node-voltage analysis method described above. Let’s briefly look at one example of another approach that can be used on its own or can simplify a complex topology before employing the node-voltage method.
  
 
[[Image:309_epd_three-R-circuit.png|thumb|right|300px|'''Three resistor circuit.''']]
 
[[Image:309_epd_three-R-circuit.png|thumb|right|300px|'''Three resistor circuit.''']]
Line 227: Line 162:
 
Yet another general circuit-solving approach is applying energy conservation equations.
 
Yet another general circuit-solving approach is applying energy conservation equations.
  
===Limits of ideal elements: more realistic model of a battery===
+
==Notes on schematic diagrams==
 
+
Recall the ideal ''i-v'' curves for our four elements: <font color=red>[Sketches repeated from first Table?]Probably just link.</font color>
+
 
+
Let’s model a battery as an ideal voltage source connected to no other elements except ideal wires.
+
 
+
<font color=red>[Sketch voltage source and wire loop.]</font color>
+
 
+
What is the current flowing through this circuit? The resistance of the wires is zero, implying that the current is infinite! Therefore this model, and the associated horizontal ''i-v'' curve, is not realistic.
+
 
+
Instead, let’s model the battery as having an internal resistance, <math>R_{BATT}</math>. To investigate the ''i-v'' curve, we’ll view the output port formed by nodes A and B.
+
 
+
<font color=red>[Sketch source, resistor, and port – include black box aspect (and in voltage divider example above) explicitly?]</font color>
+
 
+
When the port is not connected to anything, resistance is infinite and current is zero. There is no load on the battery, and thus the output voltage at node A must be the full <math>V_{BATT}</math>. This useful quantity is our by now familiar open circuit voltage, <math>v_{oc}</math>.
+
 
+
Next, let’s attach a voltmeter to the port with an associated resistance <math>R_m</math>.
+
 
+
<font color=red>[Sketch]</font color>
+
 
+
When we assume that <math>R_m = 0</math>, <math>v_A</math> is shorted to ground. Thus, the current ''i'' is at its maximum, namely the short circuit current:
+
 
+
<math>i_{sc} = \frac{V_{BATT}}{R_{BATT}}</math>.
+
 
+
These two quantities define two points on our new, more realistic, ''i-v'' curve! Presumably a line connects them, but let’s prove it by solving this simple circuit. We’ll practice our systematic approach once again.
+
 
+
The only unknown node voltage is <math>v_A</math>. The other nodes are either ground or <math>V_{BATT}</math> by definition of a perfect conductor. Hence we want to solve KCL at <math>V_{BATT}</math>, substituting in the relevant <math>v=iR</math> constitutive relations:
+
 
+
<math>i1 + i2 = 0 = \frac{9-v_A}{R_{BATT}} + \frac{0-v_A}{R_m} = 0</math> 
+
  
and substituting for <math>R_m</math> using <math>I = \frac{v_A}{R_m}</math>
+
[[Image:309_epd_shorthand-voltage-divider.png|150px|thumb|right]]
  
to simplify to <math>v_A = 9-I * R_{BATT}</math>,  
+
Another way to draw a schematic of a voltage divider circuit is shown on the right. Instead of connecting all the ground nodes together with lines, ground nodes are indicated by a ground symbol. Complex circuits have many connections to the ground node. Drawing all of the ground connections can make the schematic diagram hard to follow. This diagram represents exactly the same system as the one in which a line connected node C to the minus terminal of the voltage source.
  
indeed the equation of a line.  
+
Some other examples of topogically equivalent circuits are shown below.  
  
<font color=red>[Sketch improved battery model with slope and intercepts clearly labeled.]</font color>
+
{|
 +
|[[Image:309_epd_topological-equivalence-3R.png|250px|left|thumb|Equivalent schematic diagrams of resistors in parallel.]]
 +
|[[Image:309_epd_topological-equivalence-2R.png|250px|right|thumb|Equivalent schematic diagrams of resistors in series.]]
 +
|}
  
From the model diagram, we see that a real battery only sources a full 9V when no load is applied. As current increases, supplied voltage decreases linearly, with negative slope equalling internal resistance. <font color=red>[Need to clearly motivate why we have attached Rm! Generally, need to perturb a system to measure it… but ultimately the equation is not dependent on the added component… think further about how best to explain.]
 
  
Note that the symbol for a real battery is different from that of the ideal voltage source shown earlier: [Sketch the lines thingy versus the circle thingy.]</font color>
 
  
 
==References==
 
==References==
 
<references />
 
<references />
 +
{{Template:20.309 bottom}}

Latest revision as of 13:50, 19 October 2018

20.309: Biological Instrumentation and Measurement

ImageBar 774.jpg

Circuit diagram according to xkcd.

Overview

Resistor symbol.

The physical quantities that characterize many types of systems are defined at all points in space and time. In the case of electric circuits, the quantities of interest are charge, electric field, and magnetic field. Maxwell's equations describe the relationship between the three quantities as a function of space and time. Using Maxwell's equations to compute the field and charge at every point is a very thorough approach to analyzing electrical systems. However, such a detailed model includes far too much information to be useful for understanding or designing most circuits made from standard electronic components like resistors and capacitors. Analysis can be dramatically simplified by using a lumped element circuit model. Lumped element models characterize a system only at a small number of well-defined points by collapsing regions of space into ideal lumped elements. Regions of space that do not significantly impact system behavior are ignored.

Lumped electronic elements are idealized abstractions of physical components like resistors and capacitors. A lumped element circuit model represents a system as a network of elements connected together by wires. Each element has two state variables: current and voltage. (The lumped element model for electronic components assumes that magnetic interaction between elements are negligible.) It is possible to create lumped element models of many other types of systems described by two state variables, including thermal, acoustic, hydraulic, and mechanical systems.

Elements are bounded by terminals. In elements with two terminals, one is typically designated plus (+) and the other minus (-). Every two-terminal element has a single voltage across it and a single current flow through it. The relationship between current through the element and the voltage difference across the element's terminals is defined by a constitutive equation[1].

Each type of element has an associated graphic symbol. The symbol for a resistor is shown at right. The constitutive equation for an ideal resistor is given by Ohm's law, v=iR.

Review of electrical units and equations

Voltage

Multi-component circuit with ground nodes emphasized.

Voltage quantifies the electrical potential difference between two terminals. Higher voltages attract electrons more intensely and thus result in larger current flows. Voltage is usually represented by the variable v.

Electric potential difference has the units of energy divided by charge and is measured in volts, abbreviated with a capital V. Energy is measured in joules (J) and electric charge is measured in coulombs (C)[2]. 1 V = 1 J/C = 1 kg m2 / s2 C.

Voltage is always measured relative to a reference point. It is useful to designate a particular node as the zero point, called the ground node. You may designate any node at all as ground; however, some choices are more convenient than others. In a circuit with a single voltage or current source, the best choice is usually the negative terminal of the source. The voltage across an element in a network can be found by subtracting the node voltage at its minus terminal from the node voltage at its plus terminal.

Another possible choice of reference points is earth ground. Earth ground is the potential at which we spend most of our days — the potential of a conductor driven a few feet into the soil. If you are at earth ground potential, the doorknob doesn't shock you when you go to open the door. The ground node in most circuits is usually at a potential close to earth ground, but not always. The third conductor on electrical plugs is connected to earth ground. Since we spend most of our time near earth ground potential, connecting the outer case of electrical equipment to earth ground helps prevent shocks. Earth ground and the circuit ground node may or may not be connected together.

Current

Current measures the flow rate of charge through an element. The SI unit for current is amperes, abbreviated amps of just A. 1 A = 1 C/s. A positive current flows from the plus terminal to the minus. The variable i is usually used to represent current.

Skip the following paragraph if you want to keep things simple.

The conventional definition of current is the flow rate of positive charge carriers from the plus terminal to the minus terminal. In most of the components you will work with, such as wires, capacitors, and resistors, the charge carriers are negatively charged electrons. Electrons flowing from the minus to the plus terminal create a current with a positive sign. Current in semiconductors is carried by both positive and negative carriers. This is confusing. It's all Benjamin Franklin's fault[3]. The best approach is not to think about it at all when you are solving circuit problems. Stick to the convention that positive current flows from plus to minus. It doesn't matter for most circuits what the sign of the charge carriers is. If you are doing a gel electrophoresis, on the other hand, the sign of the charge carriers is crucial.

Resistance

Resistance has units of ohms, sometimes abbreviated with the greek letter Ω. 1 Ω = 1 V/A. Resistance is usually denoted by the letter R.

Power

Electric power is equal to voltage multiplied by current, P = v i. Power has units of energy per time called watts, abbreviated with a capital W. 1 W = 1V x 1A = 1 J / s.

Ideal lumped circuit elements

Table of ideal elements.

The table on the right summarizes the characteristics of four ideal elements: a voltage source, a current source, a resistor, and a conductor. The i-v curve in the fourth column shows the constitutive equation graphically, with voltage plotted on the vertical axis and current on the horizontal axis.

An ideal voltage source maintains a constant potential difference across its terminals regardless of the amount of current flowing through it. Thus, its voltage versus current i-v graph is a horizontal line. The current flowing through an ideal current source is constant no matter what potential difference is across its terminals. The plot is a vertical line. Ideal resistors obey Ohm’s law, v = iR, which is a sloped line on the voltage versus current graph. Ideal conductors have zero resistance. All points on a conductor have the same voltage.

You can watch a video on ideal circuit elements here.

Connecting circuit elements

Circuit illustrating nodes, branches, and loops.

Networks of elements are formed by connecting terminals together with wires. A schematic diagram like the one shown at right is one way to specify the connections. Elements interact only through their terminals. Terminals joined together by wires are called nodes. All of the terminals connected to a node have the same voltage, which is unsurprising because they are connected by an an element that by definition has the same potential everywhere. Any path between two independent nodes is called a branch. Any path from one node back around to itself is called a loop.

Terminals may be shown as filled or open circles to indicate whether or not that terminal connects to another element. Open circles represent no connection. Pairs of terminals are sometimes called ports. A port may encompass a single element or multiple elements. Dividing a circuit into groups of elements that interact through their ports can be a useful way to conceptualize and analyze systems.

Common circuit patterns

Two patterns of connected elements come up so frequently that it is worth memorizing them: series and parallel. Many circuit problems can be simplified by recognizing elements within the system that are connected in series or parallel.

Series and parallel connections

Resistors in series and parallel.

When two elements are joined end to end, they are said to be in series. In this case, the minus terminal of one element is joined to the plus terminal of the second element. The current flowing through each element connected in series connection is identical. The equivalent resistance of two resistors in series is the sum of the resistances: R = R1 + R2. Voltage sources connected in series are added.

In a parallel connection, the plus terminals of two elements are connected together, as are the minus terminals. The voltage across each of the elements in a parallel connection is the same. The equivalent resistance of two resistors in parallel is: R = R1R2/R1 + R2.

Voltage divider

309 epd basis-voltage-divider.png

A very common circuit construct consists of two resistors in series, as shown in the schematic diagram on the right. This type of circuit is called a voltage divider. In a series connection, the same current i flows through both resistors. The constitutive relation for a resistor is: v=iR. Using this equation, the voltage across resistor R1 is iR1 and the voltage across R2 is iR2. The minus terminal of R2 is connected to ground, so the voltage VC at node C is by definition 0 volts. The voltage across R2 is iR2, so the voltage VB at node B is 0+iR2=iR2. The ideal source constrains the voltage VA at node A to be V, which must also be equal to the sum of the voltages across the resistors: V = iR1+iR2=i(R1+R2). Substituting for VB into this expression gives an equation for VB in terms of the source voltage V and the resistor values: VB=VR2/R1+R2. If you don't know why the circuit is called a voltage divider, take another look at the equation.

You can watch a video on voltage dividers here.

Solving circuit problems

Charge and energy conservation

Charge and energy are conserved in electric circuits. Two equations ensure that both quantities are conserved: Kirchoff’s Current Law (KCL) and Kirchoff’s Voltage Law (KVL).

The current law states that the sum of the currents flowing into any node must equal zero. In other words, the sum of currents flowing in must equal the sum of the currents flowing out. Applying KCL to the node in-between two resistors in series demonstrates why the current is equal in both. At the shared node, current flowing out of the first element has nowhere else to go. No other branches exist at that node.

KVL ensures that energy is conserved. The law states that the sum of the potential difference across elements in a loop is zero. In other words, if you go around a loop, you end up at the same point.

KVL and KCL also offer another rationale for why the potential must be equal across elements connected in parallel. The two shared nodes form a loop. In order to sum to 0, the potential differences across the two elements must be identical.

Node-voltage analysis method

One rigorous way to solve circuit problems is to write the constitutive relation for each element (B equations for B branches), every independent KCL statement (N-1 equations for N nodes), and every independent KVL statement (B-1+N equations for B branches and N nodes). Then algebraic substitutions are made until every element voltage and current is known. Often it is not obvious which substitutions will most quickly yield the quantities that are actually of interest in that problem.

Voltage divider with current flow convention shown. Current flow in this circuit is clockwise. Hence, $ i_2 $ is negative.

The node-voltage analysis method is a faster universal approach for solving circuit problems. It indirectly incorporates KVL and requires fewer KCL statements. Rarely can there be confusion about which choice of equations will quickly yield meaningful results. (For very complex circuits, linear rather than simple algebraic approaches may be required.)

Recall that while an element voltage is the potential difference across said element, a node voltage is the potential difference between that node and a common reference point – usually ground. (If these definitions don't sound familiar, see Figure 1 under Primary definitions and also review Connecting circuit elements.)

Working with node voltages rather than element voltages makes KVL implicit, as we saw in the voltage divider example above. The basic approach is then to write a single KCL equation per unknown node voltage, while immediately substituting in the element constitutive laws in terms of voltages and resistances.[4]

To avoid mistakes when using KCL, it is helpful to adopt a consistent way of writing the node equations. One convenient way to do this is to write an expression of the current flowing into the node through each connected component, sum the expressions, and set the whole thing equal to zero. The expression for the current flowing through an element into the node is obtained by subtracting the voltage at the node under consideration from the voltage at the far end of the element (i.e. the terminal not connected to the node) and dividing by R. All you have to remember is, "the sum of the currents flowing into the node is zero."

Example: measuring voltage

Voltage divider with voltmeter attached.

Imagine connecting a voltmeter to measure the output of a voltage divider. In order to read the correct potential, the voltmeter must also be connected to ground. A model for the voltmeter includes an ideal meter and a resistance, Rm. Rm models the current i3 drawn by the voltmeter as it makes its measurement.

Before the meter is connected, KCL states that i1 + i2 = 0. Assuming R1=200Ω R2=600Ω, and V = 12 V, solving for the voltage at node B gives:


$ \frac{12-v_B}{200} + \frac{0-v_B}{600} \to v_B = 9 \text{V}. $


After the voltmeter is connected, R2 and Rm are in parallel. The KCL equation becomes: i1 + i2 + i3 = 0. If the voltmeter has a resistance of 300Ω, the voltage at node B is now:


$ \frac{12-v_B}{200} + \frac{0-v_B}{600} +\frac{0-v_B}{300} \to v_B = 6 V. $


With Rm=300Ω, the measured value is far from nominal value of 9V before the meter was connected. To make an accurate measurement, the voltmeter must have a resistance much larger than the resistors in the divider. if Rm is increased an order of magnitude to 3 kΩ, $ v_B $will be 8.57 V. Now you can see why voltmeters typically have an equivalent resistance of around 109 Ω.

Thevenin equivalent circuits

Topological equivalence between circuits implies that wires have been moved around or added in place of shorthand. There is also such a thing as functional equivalence between circuits. In fact, any complex circuit consisting only of voltage sources and resistors can be reduced to a circuit with only a single voltage source and resistor! This equivalent circuit cannot be empirically distinguished from the multi-component one.

Let’s return to the voltage divider example with a port across the second resistor. We want to know how the circuit as a whole responds at this access port when there is no load on it, i.e., no element in which energy can be dissipated. We calculate the equivalent circuit by three steps.

First we calculate the open circuit voltage ($ V_{oc} $) at that port, namely 9 V as above.

Next we calculate the short circuit current. In this hypothetical situation, the attached voltmeter has no internal resistance and behaves as a wire (Fig 12). In turn, $ v_B $ is shorted to ground, and current magnitude is dependent only on $ R_1 $, not both $ R_1 $ and $ R_2 $. Hence, the short circuit current $ I_{sc} = 12 \text{V}/200 \Omega = 0.06 \text{A} = 60 \text{mA} $.

Finally, the Thevenin resistance $ R_{Th} = \frac{V_{oc}}{I_{sc}} = \frac{9 \text{V}}{0.06 \text{A}} = 150 \Omega $. The Thevenin equivalent circuit is shown in Figure 13.

Now how does this two-component circuit reduction help us? The voltmeter (or any element!) can be connected in series with $ R_{Th} $, and $ v_B $ calculated immediately by the simple voltage divider relation, $ \frac{V R_m}{R_{tot}}. $

An alternative and direct calculation of $ R_{Th} $ is the total resistance at the port as viewed looking into the circuit (with voltage sources changed to wires): here, $ R_1 $ and $ R_2 $ in parallel.

Figure 12: Short circuit current for voltage divider.
Figure 13: Thevenin reduction of voltage divider. The Thevenin equivalent of the previously described voltage divider is shown (left) with an open port ready to be connected to a voltmeter with arbitrary resistance (right).


Alternative problem-solving strategies

Circuit problems can be approached in many ways[5] and in the interest of time we will primarily treat the approach that (nearly) universally works, namely the node-voltage analysis method described above. Let’s briefly look at one example of another approach that can be used on its own or can simplify a complex topology before employing the node-voltage method.

Three resistor circuit.

Consider the circuit shown at right. The parallel resistors $ R_2 $ and $ R_3 $ can be temporarily collapsed into an equivalent single resistor. Its resistance is given by

$ R_{parallel} = \frac{R_2 R_3}{R_2 + R_3} = 200 \Omega. $

Then, the current through the purely series-connected circuit is simply

$ i = \frac{V}{R_{tot}} = \frac{V}{R_1 + R_{parallel}} $

where $ R_{tot} = 400 \Omega $ and by the voltage divider relation $ v_A $ is clearly 6 V. That is, half the voltage drop occurs across each 200 Ω resistor, the physically real one and the imaginary one (equivalent to two real ones).

We next expand the circuit to once more include the real, parallel branches. The potential difference across each branch must be the same $ v_A $, and thus the individual currents can be calculated from $ v_A =i_{branch x}R_{branch x} $.

For example, $ i_2 =\frac{v_A}{R_2} = \frac{6}{300} = 20 \text{mA} $.

As expected, $ i_2 $ and $ i_3 $ sum to the current through the first resistor.

Yet another general circuit-solving approach is applying energy conservation equations.

Notes on schematic diagrams

309 epd shorthand-voltage-divider.png

Another way to draw a schematic of a voltage divider circuit is shown on the right. Instead of connecting all the ground nodes together with lines, ground nodes are indicated by a ground symbol. Complex circuits have many connections to the ground node. Drawing all of the ground connections can make the schematic diagram hard to follow. This diagram represents exactly the same system as the one in which a line connected node C to the minus terminal of the voltage source.

Some other examples of topogically equivalent circuits are shown below.

Equivalent schematic diagrams of resistors in parallel.
Equivalent schematic diagrams of resistors in series.


References

  1. For more about circuit assumptions and abstraction layers, see Agarwal 1.1-1.4: The power of abstraction, The lumped circuit abstraction, The lumped matter discipline, and Limitations of the lumped circuit abstraction.
  2. One mole of electrons has a charge of about 96,500 coulombs.
  3. http://www.physicsclassroom.com/class/circuits/u9l2c.cfm
  4. Slightly more complex approaches must be taken for cases in which a voltage source has a negative terminal that is not connected to ground. See Agarwal 3.3.2, The node method: Floating independent voltage sources.
  5. Overall, see chapters 2 (Resistive Networks) and 3 (Network Theorems) of Agarwal. Section 2.4 (Intuitive method of circuit analysis: series and parallel simplification) treats series/parallel reductions, section 3.5 (Superposition) treats superposition, etc.

</div>