Hydraulic analogy

Analogy between a hydraulic circuit (left) and an electronic circuit (right).

The electronichydraulic analogy (derisively referred to as the drain-pipe theory by Oliver Lodge) [1] is the most widely used analogy for "electron fluid" in a metal conductor. Since electric current is invisible and the processes at play in electronics are often difficult to demonstrate, the various electronic components are represented by hydraulic equivalents. Electricity (as well as heat) was originally understood to be a kind of fluid, and the names of certain electric quantities (such as current) are derived from hydraulic equivalents. As with all analogies, it demands an intuitive and competent understanding of the baseline paradigms (electronics and hydraulics).

Paradigms

There is no unique paradigm for establishing this analogy. Two paradigms can be used to introduce the concept to students:

Other paradigms highlight the similarities between equations governing the flow of fluid and the flow of charge:

Hydraulic analogy with horizontal water flow

Voltage, current, and charge

Hydraulic pressure difference
Electric potential 
In general, this is equivalent to hydraulic head. This model assumes that the water is flowing horizontally, so that the force of gravity can be ignored. In this case electric potential is equivalent to pressure. The voltage (or voltage drop or potential difference) is a difference in pressure between two points. Electric potential and voltage are usually measured in volts.
Current 
Equivalent to a hydraulic volume flow rate; that is, the volumetric quantity of flowing water over time. Usually measured in amperes.
Electric charge 
Equivalent to a quantity of water.

Basic circuit elements

Conducting wire
A relatively wide pipe completely filled with water is equivalent to a piece of wire. When comparing to a piece of wire, the pipe should be thought of as having semi-permanent caps on the ends. Connecting one end of a wire to a circuit is equivalent to un-capping one end of the pipe and attaching it to another pipe. With few exceptions (such as a high-voltage power source), a wire with only one end attached to a circuit will do nothing; the pipe remains capped on the free end, and thus adds nothing to the circuit.
Resistor 
A constriction in the bore of the pipe which requires more pressure to pass the same amount of water. All pipes have some resistance to flow, just as all wires have some resistance to current.
Node (or junction) in Kirchhoff's junction rule 
A pipe tee. The net flow of water into a piping tee (filled with water) must equal the net flow out.
Capacitor 
A tank with one connection at each end and a rubber sheet dividing the tank in two lengthwise[7] (a hydraulic accumulator). When water is forced into one pipe, equal water is simultaneously forced out the other pipe, yet no water can penetrate the rubber diaphragm. Energy is stored by the stretching of the rubber. As more current flows "through" the capacitor, the back-pressure (voltage) becomes greater, thus current "leads" voltage in a capacitor. As the back-pressure from the stretched rubber approaches the applied pressure, the current becomes less and less. Thus capacitors "filter out" constant pressure differences and slowly varying, low-frequency pressure differences, while allowing rapid changes in pressure to pass through.
Inductor 
A heavy paddle wheel placed in the current. The mass of the wheel and the size of the blades restrict the water's ability to rapidly change its rate of flow (current) through the wheel due to the effects of inertia, but, given time, a constant flowing stream will pass mostly unimpeded through the wheel, as it turns at the same speed as the water flow. The mass and surface area of the wheel and its blades are analogous to inductance, and friction between its axle and the axle bearings corresponds to the resistance that accompanies any non-superconducting inductor.
An alternative inductor model is simply a long pipe, perhaps coiled into a spiral for convenience. This fluid-inertia device is used in real life as an essential component of a hydraulic ram. The inertia of the water flowing through the pipe produces the inductance effect; inductors "filter out" rapid changes in flow, while allowing slow variations in current to be passed through. The drag imposed by the walls of the pipe is somewhat analogous to parasitic resistance.
In either model, the pressure difference (voltage) across the device must be present before the current will start moving, thus in inductors voltage "leads" current. As the current increases, approaching the limits imposed by its own internal friction and of the current that the rest of the circuit can provide, the pressure drop across the device becomes lower and lower.
Ideal voltage source (ideal battery) or ideal current source 
A dynamic pump with feedback control. A pressure meter on both sides shows that regardless of the current being produced, this kind of pump produces constant pressure difference. If one terminal is kept fixed at ground, another analogy is a large body of water at a high elevation, sufficiently large that the drawn water does not affect the water level. To create the analog of an ideal current source, use a positive displacement pump: A current meter (little paddle wheel) shows that when this kind of pump is driven at a constant speed, it maintains a constant speed of the little paddle wheel.

Other circuit elements

Diode 
Equivalent to a one-way check valve with a slightly leaky valve seat. As with a diode, a small pressure difference is needed before the valve opens. And like a diode, too much reverse bias can damage or destroy the valve assembly.
Transistor 
A valve in which a diaphragm, controlled by a low-current signal (either constant current for a BJT or constant pressure for a FET), moves a plunger which affects the current through another section of pipe.
CMOS 
A combination of two MOSFET transistors. As the input pressure changes, the pistons allow the output to connect to either zero or positive pressure.
Memristor 
A needle valve operated by a flow meter. As water flows through in the forward direction, the needle valve restricts flow more; as water flows the other direction, the needle valve opens further providing less resistance.

Principle equivalents

EM wave speed (velocity of propagation) 
Speed of sound in water. When a light switch is flipped, the electric wave travels very quickly through the wires.
Charge flow speed (drift velocity) 
Particle speed of water. The moving charges themselves move rather slowly.
DC 
Constant flow of water in a circuit of pipe
Low frequency AC 
Water oscillating back and forth in a pipe
Higher-frequency AC and transmission lines 
Sound being transmitted through the water pipes (this does not properly mirror the cyclical reversal of alternating electric current). As described, the fluid flow conveys pressure fluctuations, but fluids do not reverse at high rates in hydraulic systems, which the above "low frequency" entry does accurately describe. A better concept (if sound waves are to be the phenomenon) is that of direct current with high-frequency "ripple" superimposed.
Inductive spark 
Used in induction coils, similar to water hammer, caused by the inertia of water

See also Bond graph

Equation examples

Some examples of equivalent electrical and hydraulic equations:

type hydraulic electric thermal mechanical
quantity volume [m3] charge [C] heat [J] momentum [Ns]
potential pressure [Pa=J/m3=N/m2] potential [V=J/C=W/A] temperature [K=J/] velocity [m/s]
flux Volumetric flow rate [m3/s] current [A=C/s] heat transfer rate [J/s] force [N]
flux density velocity [m/s] current density [C/(m2·s) = A/m²] heat flux [W/m2] stress [N/m2 = Pa]
linear model Poiseuille's law Ohm's law Fourier's law Dashpot

If the differential equations have the same form, the response will be similar.

Limits to the analogy

If taken too far, the water analogy can create misconceptions. For it to be useful, one must remain aware of the regions where electricity and water behave very differently.

Fields (Maxwell equations, Inductance)
Electrons can push or pull other distant electrons via their fields, while water molecules experience forces only from direct contact with other molecules. For this reason, waves in water travel at the speed of sound, but waves in a sea of charge will travel much faster as the forces from one electron are applied to many distant electrons and not to only the neighbors in direct contact. In a hydraulic transmission line, the energy flows as mechanical waves through the water, but in an electric transmission line the energy flows as fields in the space surrounding the wires, and does not flow inside the metal. Also, an accelerating electron will drag its neighbors along while attracting them, both because of magnetic forces.
Charge 
Unlike water, movable charge carriers can be positive or negative, and conductors can exhibit an overall positive or negative net charge. The mobile carriers in electric currents are usually electrons, but sometimes they are charged positively, such as the positive ions in an electrolyte, the H+ ions in proton conductors or holes in p-type semiconductors and some (very rare) conductors.
Leaking pipes
The electric charge of an electrical circuit and its elements is usually almost equal to zero, hence it is (almost) constant. This is formalized in Kirchhoff's current law, which does not have an analogy to hydraulic systems, where amount of the liquid is not usually constant. Even with incompressible liquid the system may contain such elements as pistons and open pools, so the volume of liquid contained in a part of the system can change. For this reason, continuing electric currents require closed loops rather than hydraulics' open source/sink resembling spigots and buckets.
Fluid velocity and resistance of metals 
As with water hoses, the carrier drift velocity in conductors is directly proportional to current. However, water only experiences drag via the pipes' inner surface, while charges are slowed at all points within a metal, as with water forced through a filter. Also, typical velocity of charge carriers within a conductor is less than centimeters per minute, and the "electrical friction" is extremely high. If charges ever flowed as fast as water can flow in pipes, the electric current would be immense, and the conductors would become incandescently hot and perhaps vaporize. To model the resistance and the charge-velocity of metals, perhaps a pipe packed with sponge, or a narrow straw filled with syrup, would be a better analogy than a large-diameter water pipe. Resistance in most electrical conductors is a linear function: as current increases, voltage drop increases proportionally (Ohm's Law). Liquid resistance in pipes is not linear with volume, varying as the square of volumetric flow (see Darcy–Weisbach equation).
Quantum Mechanics 
Solid conductors and insulators contain charges at more than one discrete level of atomic orbit energy, while the water in one region of a pipe can only have a single value of pressure. For this reason there is no hydraulic explanation for such things as a battery's charge pumping ability, a diode's depletion layer and voltage drop, solar cell functions, Peltier effect, etc., however equivalent devices can be designed which exhibit similar responses, although some of the mechanisms would only serve to regulate the flow curves rather than to contribute to the component's primary function.

Usefulness requires that the reader or student has a substantial understanding of the model (hydraulic) system's principles. It also requires that the principles can be transferred to the target (electrical) system. Hydraulic systems are deceptively simple: the phenomenon of pump cavitation is a known, complex problem that few people outside of the fluid power or irrigation industries would understand. For those who do, the hydraulic analogy is amusing, as no "cavitation" equivalent exists in electrical engineering. The hydraulic analogy can give a mistaken sense of understanding that will be exposed once a detailed description of electrical circuit theory is required.

One must also consider the difficulties in trying to make an analogy match reality completely. The above "electrical friction" example, where the hydraulic analog is a pipe filled with sponge material, illustrates the problem: the model must be increased in complexity beyond any realistic scenario.

See also

Notes

  1. Paul J. Nahin, Oliver Heaviside: The Life, Work, and Times of an Electrical Genius of the Victorian Age, JHU Press, 2002 ISBN 0801869099 page 59
  2. A. Akers, M. Gassman, & R. Smith, Hydraulic Power System Analysis. Taylor & Francis, New York, 2006, Chapter 13, ISBN 0-8247-9956-9.
  3. A. Esposito, "A Simplified Method for Analyzing Circuits by Analogy". Machine Design, October 1969, pp. 173-177.
  4. Brian J. Kirby, Micro- and Nanoscale Fluid Mechanics, p. 69, Cambridge University Press, 2010 ISBN 1139489836.
  5. Schelleng, John C. "The violin as a circuit." The Journal of the Acoustical Society of America 35.3 (2005): 326-338. http://www.maestronet.com/forum/index.php?app=core&module=attach&section=attach&attach_id=13435
  6. "Axial feedback stabilization of a flute mode in a simple mirror reactor, by M. A. Lieberman and S. L. Wong, Plasma Physics, Vol. 19, pp. 745-55 (1977). The article contains an L-C circuit that is unstable because the "capacitance" is negative: http://iopscience.iop.org/0032-1028/19/8/005/pdf/0032-1028_19_8_005.pdf
  7. http://amasci.com/emotor/cap1.html
Wikimedia Commons has media related to Hydraulic analogy.
This article is issued from Wikipedia - version of the 9/20/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.