You are on page 1of 211

An Introduction to Waterpower

Water power has been harnessed since the earliest civilizations for agricultural processing and latterly
for the production of electricity. If you live in an old mill or have an upland stream flowing through your
property, you probably have a site suitable to generate all or a significant proportion of your domestic
power requirements.
If there is insufficient power to provide your needs electrically, then there is the additional possibility of
using water power to drive a water source heat pump which will increase the power output by three
times in the form of heat.
The main power requirements are in the winter months, so there should be little or no conflict with
other river demands. Adequate provisions may have to be made for migratory fish, and if new works are
to be constructed, care must be taken with regard to the effect on land drainage upstream of your
intake. We recommend that our customers construct all schemes in an unobtrusive and
environmentally sensitive manner, and where appropriate we will provide drawings of proven designs.
Available Power
The power available from a stream is determined by the head and flow of water on the particular site.
This power is harnessed by constructing a dam or diverting the flow in such a way that all the fall occurs
in one place. Where it is not practical to construct a channel, the water may be piped and the head of
water is exploited as a high velocity jet driving an Impulse Turbine.
The power available is a function of the fall (head) and flow so building a large waterwheel on a low fall
will only increase the cost and reduce the shaft speed but not increase the power.
Water wheels are limited to sites with a head of less than 10 meters. They are aesthetically pleasing and
have good performance under low water conditions. Unfortunately, due to their size, they are both
costly to build and install, largely because of the gearing required to increase the shaft speed, typically
from 10 to 1500 rpm. The use of low speed generators does not help since it is the low speed end of the
drive which is the expensive part.
Water turbines, on the other hand, are able to make use of a very wide range of head, from less than a
metre to many hundreds of metres. To cover the full range of sites, it is necessary to make use of
several different types of turbine. It is not that you cannot use one type of turbine for all sites but that
each design has it economic and hydraulic area.
Types of waterpower site
Upland sites
Small streams typically employ a pipeline of 100 mm to 600 mm in diameter running downhill to the
power plant which may have an operating head in excess of 20 metres. The pipeline could be as long as
1000 metres and still be economic, provided the gradient is steep enough, or most of the pipe run can
utilize low cost materials. An impulse turbine having one or more jets is typical for this type of
installation.
Low head sites
These use proportionately greater quantities of water frequently employ an open leat or channel to
bring the water around the contour to a point where it can drop steeply down to the turbine or
millhouse. Such sites with heads in excess of 10 metres can also use impulse turbines which can adapt
well to varying water flows that occur between summer and winter.
For thousands of years waterpower has been harnessed for milling and pumping water but many good
sites have fallen into disrepair as a result of competition from diesel and electric power Of over 70,000

mills in the UK, at the end of the 18th century, there are only a few hundred still working. Mills fall into
several categories which will determine their suitability for redevelopment.
Mills with ponds
Seldom suitable for redevelopment for anything other than a few kilowatts because the water flow is
obviously too little to sustain the mill on a continuous basis, and it is usually too expensive to install a
wheel or turbine that can only be operated for a few hours a day.
In some cases the ponds were only needed in the summer months when the water flow was low and
this can still be a useful system where the equipment to be driven requires a fixed power input. In such
situations the plant operates until the water level is reduced by say 500mm, it is then stopped until the
pond fills again.
Although it is not always very convenient, it is efficient, since the plant operates at full power when
turned on. This system is suitable for heating, pumping and where there is an alternative source of
power if it is required all the time.
Mills with leats or channels
These take their water from a water course along the side of a valley at a gradient that is usually less
than one in five hundred. At a suitable point when enough fall can be achieved in one place.
If a turbine is installed to replace a waterwheel, it may be possible to increase the useful head by
eliminating drops at mill pools, sluices and below the wheel. Since there was a mill there anyway there
should be at least enough power for domestic purposes including heating. Improvements to the leat and
head are usually possible but are very site specific and are made very much easier with mini excavators.
Mills on weirs
Weirs or lcations with short wide diversion channels present the most difficult challenge for the
developer. The available head may only be a metre or so and the flow required to generate useful
amounts of power will be several cubic metres of water per second.
The undershot waterwheels that were originally used at these sites are totally redundant on account of
their high cost and low efficiency. The exact layout of the site becomes increasingly important with the
lower falls, because access for excavators and to install the large items of equipment is more difficult.
Water falls ranging from one to three metres in height employ single or multiple high speed propeller
turbines. Variable flow is achieved by automatically stopping and starting one or more of the turbines
rather than using variable flow Kaplan turbines that are more expensive and less efficient.
Very low head and tidal causeway
These sites can be developed using modular bulb turbines. Propeller turbines, driving high pressure
water pumps are arranged in a line across the river and the water is fed to another impulse turbine on
the river bank. The lower efficiency is compensated for by lower maintenance costs and variable speed
capability for the main turbines. Our largest units to date, have a runner diameter of 3.5 metres.
The range of project size
Applications range from power for remote monitoring equipment, telemetry or camp lighting, to small
scale village electrification in remote rural areas. Our Universal range of turbines has been designed to
offer an alternative to diesel plants and to have much more flexibility in their installation than
conventional waterpower equipment.
Larger projects over 100 kW for commercial power generation, need to be individually designed to
make the best use of the power available Special projects have included very small units for battery
charging, power recovery in industry and low-cost units for agricultural applications and water pumping.

Developing Countries Simple robust technology is required to drive agricultural machinery; typically
supplied bare shaft with mechanical governing. The machines are suitable for local manufacture.
Portable plants (5-15kW) For rapid installation and relocation for disaster relief operations, survey
camp power and recreation purposes.
Domestic power (5-25kW) Supply for one house if heating is the main load, usually single phase. Sites
include old mills and upland farms with or without grid connection.
Commercial power (15-100kW) Usually three phase installations supplying small businesses, farms or
estates. Can run in parallel with the grid and have accurate governing systems.
Grid connected plants (50-2500kW) Specifically designed for commercial power connected to grid
system. Designed to run fully automated with remote monitoring.
Training purposes
Models for training purposes either static for illustration only or as fully working desk top units,
complete with pump tank and generator.
Special applications
Outside normal power and head range and are individually assessed. These include power recovery in
desalination or industrial processes, aeration for fish farms, sub-sea power, fire fighting, mining, tidal
and canal lock installations.
Power recovery
Turbines to replace head breaking valves or at the inlet to treatment works; for desalination plants using
reverse osmosis; for compensation flows from reservoirs and the outfalls of sewage treatment facilities.
In line power recovery on contaminated flows and raw sewage is also possible
Micro hydropower Environmentally friendly Micro hydro, or small-scale hydro, is one of the most
environmentally benign energy conversion options available, because unlike large-scale hydro power, it
does not attempt to interfere significantly with river flows.
When is hydropower micro? The definition of micro hydropower varies in different countries and can
even include systems with a capacity of a few megawatts. One of the many definitions for micro
hydropower is: hydro systems up to a rated capacity of approximately 300 kW capacity. The limit is set
to 300 kW because this is about the maximum size for most stand alone hydro systems not connected
to the grid, and suitable for "run-of-the-river" installations.
Microhydro definitions Definitions of small, mini, and micro hydro plants found in litreature
(kilowatts)

mini
small
(kilowatts) (megawatts)

United States < 100

100 - 1000

1 - 30

United States < 100

100 - 1000

China

< 500

0.5 - 25

USSR

< 100

0.1 - 30

France

5 - 5000

India

< 100

101 - 1000

1 -15

Brazil

< 100

100 - 1000

1 - 30

Norway *

< 100

100 - 1000

1 - 10

various

< 100

< 1000

< 10

Small Michell (Banki) Turbine:


A CONSTRUCTION MANUAL
I. WHAT IT IS AND HOW IT IS USEFUL
The Michell or Banki turbine is a relatively easy to build and highly efficient means of harnessing a small
stream to provide enough power to generate electricity or drive different types of mechanical devices.
FIGURE 1

The turbine consists of two main parts--the runner, or wheel, and the nozzle. Curved horizontal blades
are fixed between the circular end plates of the runner (see page 17). Water passes from the nozzle
through the runner twice in a narrow jet before it is discharged.
Once the flow and head of the water site have been calculated, the blades of the 30cm diameter wheel
presented here can be lengthened as necessary to obtain optimum power output from the available
water source.
The efficiency of the Michell turbine is 80 percent or greater. This, along with its adaptability to a variety
of water sites and power needs, and its simplicity and low cost, make it very suitable for small power
development.
The turbine itself provides power for direct current (DC); a governing device is necessary to provide
alternating current (AC).
II. DECISION FACTORS
Applications:
* Electric generation (AC or DC)
* Machinery operations, such as threshers, winnower, water pumping, etc.
Advantages:
* Very efficient and simple to build and operate.

* Virtually no maintenance.
* Can operate over a range of water flow and head conditions.
Considerations:
* Requires a certain amount of skill in working with metal.
* Special governing device is needed for AC electric generation.
* Welding equipment with cutting attachments are needed.
* Electric grinding machine is needed.
Access to small machine shop is necessary.
COST ESTIMATE(*)
$150 to $600 (US, 1979) including materials and labour.
(This is for the turbine only. Planning and construction costs of dam, penstock, etc., must be added.)
(*) Cost estimates serve only as a guide and will vary from country to country.
PLANNING
Development of small water power sites currently comprises one of the most promising applications of
alternate energy technologies. If water power will be used to produce only mechanical energy--for
example, for powering a grain thresher--it may be easier and less expensive to construct a waterwheel
or a windmill. However, if electrical generation is needed, the Michell turbine, despite relatively high
initial costs, may be feasible and indeed economical under one or more of the following conditions:
* Access to transmission lines or to reliable fossil fuel sources is limited or non-existent.
* Cost of fossil and other fuels is high.
* Available water supply is constant and reliable, with a head of 50-100m relatively easy to achieve.
* Need exists for only a small dam built into a river or stream and for a relatively short (less than 35m)
penstock (channel) for conducting water to the turbine.
If one or more of the above seems to be the case, it is a good idea to look further into the potential of a
Michell turbine. The final decision will require consideration of a combination of factors, including site
potential, expense, and purpose.
III. MAKING THE DECISION AND FOLLOWING THROUGH
When determining whether a project is worth the time, effort, and expense involved, consider social,
cultural, and environmental factors as well as economic ones. What is the purpose of the effort? Who
will benefit most? What will the consequences be if the effort is successful? And if it fails?
Having made an informed technology choice, it is important to keep good records. It is helpful from the
beginning to keep data on needs, site selection, resource availability, construction progress, labour and
materials costs, test findings, etc. The information may prove an important reference if existing plans
and methods need to be altered. It can be helpful in pinpointing "what went wrong?" And, of course, it
is important to share data with other people.
The technologies presented in this and the other manuals in the energy series have been tested
carefully and are actually used in many parts of the world. However, extensive and controlled field tests
have not been conducted for many of them, even some of the most common ones. Even though we
know that these technologies work well in some situations, it is important to gather specific information
on why they perform properly in one place and not in another.
Well documented models of field activities provide important information for the development worker.
It is obviously important for a development worker in Colombia to have the technical design for a
machine built and used in Senegal. But it is even more important to have a full narrative about the
machine that provides details on materials, labour, design changes, and so forth. This model can provide
a useful frame of reference.
A reliable bank of such field information is now growing. It exists to help spread the word about these
and other technologies, lessening the dependence of the developing world on expensive and finite
energy resources.
A practical record keeping format can be found in Appendix IV.
IV. PRE-CONSTRUCTION CONSIDERATIONS

Both main parts of the Michell turbine are made of plate steel and require some machining. Ordinary
steel pipe is cut to form the blades or buckets of the runner. Access to welding equipment and a small
machine shop is necessary.
The design of the turbine avoids the need for a complicated and well-sealed housing. The bearings have
no contact with the water flow, as they are located outside of the housing; they can simply be lubricated
and don't need to be sealed.
Figure 2 shows an arrangement of a turbine of this type for low-head use without control. This
installation will drive an AC or DC generator with a belt drive.

SITE SELECTION
This is a very important factor. The amount of power obtained, the expense of installation, and even, by
extension, the applications for which the power can be used may be determined by the quality of the
site.
The first site consideration is ownership. Installation of an electricity-generating unit--for example, one
that needs a dam and reservoir in addition to the site for the housing--can require access to large
amounts of land.
In many developing countries, large lots of land are few and it is likely that more than one owner will
have to be consulted. If ownership is not already clearly held, the property questions must be
investigated, including any rights which may belong to those whose property borders on the water.
Damming, for example, can change the natural water flow and/or water usage patterns in the area and
is a step to be taken only after careful consideration.
If ownership is clear, or not a problem, a careful analysis of the site is necessary in order to determine:
1) the feasibility of the site for use of any kind, and 2) the amount of power obtainable from the site.
Site analysis consists of collecting the following basic data:
* Minimum flow.
* Maximum flow.
* Available head (the height a body of water falls before hitting the machine).

* Pipe line length (length of penstock required to give desired head).


* Water condition (clear, muddy, sandy, acid, etc.).
* Site sketch (with evaluations, or topographical map with site sketched in).
* Soil condition (the size of the ditch and the condition of the soil combine to affect the speed at which
the water moves through the channel and, therefore, the amount of power available).
* Minimum tailwater (determines the turbine setting and type).
Appendix I contains more detailed information and the instructions needed to complete the site analysis
including directions for measuring head, water flow, and head losses. These directions are simple
enough to be carried out in field conditions without a great deal of complex equipment.
Once such information is collected, the power potential can be determined. Some power, expressed in
terms of horsepower or kilowatts (one horsepower equals 0.7455 kilowatts), will be lost because of
turbine and generator inefficiencies and when it is transmitted from the generator to the place of
application.
For a small water power installation of the type considered here, it is safe to assume that the net power
(power actually delivered) will be only half of the potential gross power.
Gross power, or power available directly from the water, is determined by the following formula:
Gross Power
Gross power (English units: horsepower) =
Minimum Water Flow (cubic feet/second) X Gross Head (feet)
8.8
Gross power (metric horsepower) =
1,000 Flow (cubic meters/second) X Head (meters)
75
Net Power (available at the turbine shaft)
Net Power (English units) =
Minimum Water Flow X Net Head(*) X Turbine Efficiency
8.8
Net Power (metric units) =
Minimum Water Flow X Net Head(*) X Turbine Efficiency
75/1,000
(*) Net head is obtained by deducting energy losses from the gross head (see page 57).
A good assumption for turbine efficiency when calculating losses is 80 percent.
Some sites lend themselves naturally to the production of electrical or mechanical power. Other sites
can be used if work is done to make them suitable. For example, a dam can be built to direct water into
a channel intake or to get a higher head than the stream provides naturally. (A dam may not be required
if there is sufficient head or if there is enough water to cover the intake of a pipe or channel leading to
the penstock.) Dams may be of earth, wood, concrete, or stone.
Appendix II provides some information on construction of small dams.
EXPENSE
Flowing water tends to generate automatically a picture of "free" power in the eyes of the observer. But
there is always a cost to producing power from water sources. Before proceeding, the cost of
developing low-output water power sites should be checked against the costs of other possible
alternatives, such as:
* Electric utility--In areas where transmission lines can furnish unlimited amounts of reasonably priced
electric current, it is often uneconomical to develop small or medium-sized sites.
However, in view of the increasing cost of utility supplied electricity, hydroelectric power is becoming
more cost-effective.
* Generators--Diesel engines and internal-combustion engines are available in a wide variety of sizes
and use a variety of fuels--for example, oil, gasoline, or wood. In general, the capital expenditure for this

type of power plant is low compared to a hydroelectric plant. Operating costs, on the other hand, are
very low for hydroelectric and high for fossil fuel generated power.
* Solar--Extensive work has been done on the utilization of solar energy for such things as water
pumping. Equipment now available may be less costly than water power development in regions with
long hours of intense sunshine.
If it seems to make sense to pursue development of the small water power site, it is necessary to
calculate in detail whether the site will indeed yield enough power for the specific purposes planned.
Some sites will require investing a great deal more money than others. Construction of dams and
penstocks can be very expensive, depending upon the size and type of dam and the length of the
channel required. Add to these construction expenses, the cost of the electric equipment--generators,
transformers, transmission lines--and related costs for operation and maintenance and the cost can be
substantial.
Any discussion of site or cost, however, must be done in light of the purpose for which the power is
desired. It may be possible to justify the expense for one purpose but not for another.
ALTERNATING OR DIRECT CURRENT
A turbine can produce both alternating (AC) and direct current (DC). Both types of current cannot
always be used for the same purposes and one requires installation of more expensive equipment than
the other.
Several factors must be considered in deciding whether to install an alternating or direct current power
unit.
The demand for power will probably vary from time to time during the day. With a constant flow of
water into the turbine, the power output will thus sometimes exceed the demand.
In producing AC, either the flow of water or the voltage must be regulated because AC cannot be stored.
Either type of regulation requires additional equipment which can add substantially to the cost of the
installation.
The flow of water to a DC-producing turbine, however, does not have to be regulated. Excess power can
be stored in storage batteries. Direct current generators and storage batteries are relatively low in cost
because they are mass-produced.
Direct current is just as good as AC for producing electric light and heat. But electrical equipment having
AC motors, such as farm machinery and household appliances, have to be changed to DC motors. The
cost of converting appliances must be weighed against the cost of flow regulation needed for producing
AC.
APPLICATIONS
While a 30.5cm diameter wheel has been chosen for this manual because this size is easy to fabricate
and weld, the Michell turbine has a wide range of application for all water power sites providing head
and flow are suitable. The amount of water to be run through the turbine determines the width of the
nozzle and the width of the wheel. These widths may vary from 5cm to 36cm. No other turbine is
adaptable to as large a range of water flow (see Table 1).
Impulse or Pelton Michell or Banki Centrifugal Pump Used as Turbine
Head Range (feet ): 50 to 1000 3 to 650
Flow Range (cubic) feet per second: 0.1 to 10 0.5 to 250
Application: high head, medium head. Available for any desired condition
Power (horsepower): 1 to 500 1 to 1000
Cost per Kilowatt: low
Table 1. Small Hydraulic Turbines
The size of the turbine depends on the amount of power required, whether electrical or mechanical.
Many factors must be considered to determine what size turbine is necessary to do the job. The
following example illustrates the decision-making process for the use of a turbine to drive a peanut
huller (see Figure 3).

Steps will be similar in electrical power applications.


* Power enough to replace the motor for a 2-1/2 hp 1800 revolutions per minute (rpm) peanut
thresher.
* Gross power needed is about 5 hp (roughly twice the horsepower of the motor to be replaced
assuming that the losses are about one-half of the total power available).
* Village stream can be dammed up and the water channeled through a ditch 30m (100 ft) long.
* Total difference in elevation is 7.5m (25 ft).
* Available minimum flow rate: 2.8 cu ft/sec.
* Soil of ditch permits a water velocity of 2.4 ft/sec (Appendix I, Table 2 gives n = 0.030).
* Area of flow in ditch = 2.8/2.4 - 1.2 sq ft.
* Bottom width = 1.2 ft.
* Hydraulic radius = 0.31 x 1.2 = 0.37 ft (see Appendix I).
Calculate results of fall and head loss.
Shown on nomograph (Appendix I) as a 1.7 foot loss for every 1,000 feet. Therefore the total loss for a
30m (100 ft) ditch is:
1.7/10 = 0.17 feet
Since 0.17 ft is a negligible loss, calculate head at 25 ft.
Power produced by turbine at 80% efficiency = 6.36 hp
Net power = Minimum water flow x net head x turbine efficiency
8.8
2.8 x 25 x 0.80
8.8 = 6.36 horsepower
Formulas for principal Michell turbine dimensions:
([B.sub.1]) = width of nozzle = 210 x flow
-------------------------------------------Runner outside diameter x [square root] head
=
210 x 2.8 = 9.8 inches
---------------------12 x [square root] 25
([B.sub.2]) = width of runner between discs - ([B.sub.1]) = 1/2 to 1 inch
= 9.8 + 1 inch = 10.8 inches
Rotational speed (revolutions per minute)
=
73.1 x [square root] head
---------------------------Runner outside diameter (ft)
73.1 x [square root] 25
----------------------1
= 365.6 rpm

The horsepower generated is more than enough for the peanut huller but the rpm is not high enough.
Many peanut threshers will operate at varying speeds with proportional yield of hulled peanuts. So for a
huller which gives maximum output at 2-1/2 hp and 1800 rpm, a pulley arrangement will be needed for
stepping up speed. In this example, the pulley ratio needed to step up speed is 1800 .365 or
approximately 5:1. Therefore a 15" pulley attached to the turbine shaft, driving a 3" pulley on a
generator shaft, will give [+ or -] 1800 rpm.
MATERIALS
Although materials used in construction can be purchased new, many of these materials can be found at
junk yards.
Materials for 30.5cm diameter Michell turbine:
* Steel plate 6.5mm X 50cm X 100cm
* Steel plate 6.5mm thick (quantity of material depends on nozzle width)
* 10cm ID water pipe for turbine buckets(*)
* Chicken wire (1.5cm X 1.5cm weave) or 25mm dia steel rods
* 4 hub flanges for attaching end pieces to steel shaft (found on most car axles)
* 4.5cm dia solid steel rod
* two 4.5cm dia pillow or bush bearings for high speed use. (It is possible to fabricate wooden bearings.
Because of the high speed, such bearings would not last and are not recommended.)
* eight nuts and bolts, appropriate size for hub flanges
TOOLS
* Welding equipment with cutting attachments
* Metal file
* Electric or manual grinder
* Drill and metal bits
* Compass and Protractor
* T-square (template included in the back of this manual)
* Hammer
* C-clamps
* Work bench
(*) Measurements for length of the pipe depend on water site conditions.
V. CONSTRUCTION
PREPARE THE END PIECES
An actual size template for a 30.5cm turbine is provided at the end of this manual. Two of the bucket
slots are shaded to show how the buckets are installed.
Figure 4 shows the details of a Michell runner.

* Cut out the half circle from the template and mount it on cardboard or heavy paper.
* Trace around the half circle on the steel plate as shown in Figure 5.

* Turn the template over and trace again to complete a full circle (see Figure 6.

* Draw the bucket slots on the template with a clockwise slant as shown in Figure 7.

* Cut out the bucket slots on the template so that there are 10 spaces.
* Place the template on the steel plate and trace in the bucket slots.
* Repeat the tracing process as before to fill in the area for the shaft (see Figure 8).

* Drill a 2mm hole in the steel plate in the center of the wheel where the cross is formed. The hole will
serve as a guide for cutting the metal plate.
FIGURE 9

* Take a piece of scrap metal 20cm long x 5cm wide. Drill a hole the width of the opening in the torch
near one end of the metal strip.
* Drill a 2mm dia hole at the other end at a point equal to the radius of the wheel (15.25cm). Measure
carefully.
* Line up the 2mm hole in the scrap metal with the 2mm hole in the metal plate and attach with a nail
as shown in Figure 10.

* Cut both end plates as shown (in Figure 10) using the torch.
* Cut the bucket slots with the torch or a metal saw.
* Cut out a 4.5cm dia circle from the center of both wheels. This prepares them for the axle.
CONSTRUCT THE BUCKETS
Calculate the length of buckets using the following formula:
Width of Buckets = 210 x Flow (cu/ft/sec)
+ (1 .5in) Between End Plates Outside Diameter of Turbine (in) x [square root] Head (ft)
* Once the bucket length has been determined, cut the 10cm dia pipe to the required lengths.
* When cutting pipe lengthwise with a torch, use a piece of angle iron to serve as a guide, as shown in
Figure 11.

(Bucket measurements given in the template in the back of this manual will serve as a guide.)
* Pipe may also be cut using an electric circular saw with a metal cutting blade.
* Cut four buckets from each section of pipe. A fifth piece of pipe will be left over but it will not be the
correct width or angle for use as a bucket (see Figure 12).

* File each of the buckets to measure 63mm wide. (NOTE: Cutting with a torch may warp the buckets.
Use a hammer to straighten out any warps.)
ASSEMBLE THE TURBINE
* Cut a shaft from 4.5cm dia steel rod. The total length of the shaft should be 60cm plus the width of the
turbine.
* Place the metal hubs on the center of each end piece, matching the hole of the hub with the hole of
the end piece.
* Drill four 20mm holes through the hub and end piece.
* Attach a hub to each end piece using 20mm dia x 3cm long bolts and nuts.
* Slide shaft through the hubs and space the end pieces to fit the buckets.
FIGURE 13

* Make certain the distance from each end piece to the end of the shaft is 30cm.
* Insert a bucket and align the end pieces so that the blade runs perfectly parallel with the center shaft.
* Spot weld the bucket in place from the outside of the end piece (see Figure 14).

* Turn the turbine on the shaft half a revolution and insert another bucket making sure it is aligned with
the center shaft.
* Spot weld the second bucket to the end pieces. Once these buckets are placed, it is easier to make
sure that all the buckets will be aligned parallel to the center shaft.
* Weld the hubs to the shaft (check measurements).
* Weld the remaining buckets to the end pieces (see Figure 15).

* Mount the turbine on its bearings. Clamp each bearing to the workbench so that the whole thing can
be slowly rotated as in a lathe. The cutting tool is an electric or small portable hand grinder mounted on
a rail and allowed to slide along a second rail, or guide (see Figure 16). The slide rail should be carefully
clamped so that it is exactly parallel to the turbine shaft.

* Grind away any uneven edges or joints. Rotate the turbine slowly so that the high part of each blade
comes into contact with the grinder. Low parts will not quite touch. This process takes several hours and
must be done carefully.
* Make sure the bucket blades are ground so that the edges are flush with the outside of the end pieces.
* Balance the turbine so it will turn evenly (see Figure 17).

It may be necessary to weld a couple of small metal washers on the top of either end of the turbine. The
turbine is balanced when it can be rotated in any position without rolling.
MAKE THE TURBINE NOZZLE
* Determine nozzle size by using the following formula:
210 X flow (cubic feet/second
-----------------------------------------------------runner outside diameter (in) x [square root] head (ft)
The nozzle should be 1.5cm to 3cm less than the inside width of the turbine.
Figure 18 shows a front view of a properly positioned nozzle in relationship to the turbine.

* From a 6.5mm steel plate, cut side sections and flat front and back sections of the nozzle. Width of
front and back pieces will be equal to the width of the turbine wheel minus 1.5 to 3cm. Determine other
dimensions from the full-scale diagram in Figure 19.

* Cut curved sections of the nozzle from 15cm (OD) steel pipe if available. Make sure that the pipe is
first cut to the correct width of the nozzle as calculated previously. (Bend steel plate to the necessary
curvature if 15cm pipe is unavailable. The process will take some time and ingenuity on the part of the
builder. One way of bending steel plate is to sledge hammer the plate around a steel cylinder or
hardwood log 15cm in diameter. This may be the only way to construct the nozzle if 15cm steel pipe is
unavailable.)
* Weld all sections together. Follow assembly instructions given in "Turbine Housing".
The diagram in Figure 19 provides minimum dimensions for proper turbine installation.
TURBINE HOUSING
Build the structure to house the turbine and nozzle of concrete, wood, or steel plate. Figure 20 shows a
side view and front view of a typical installation for low head use (1-3m). Be sure housing allows for easy
access to the turbine for repair and maintenance.

* Attach the nozzle to the housing first and then orient the turbine to the nozzle according to the
dimensions given in the diagram in Figure 19. This should ensure correct turbine placement. Mark the
housing for the placement of the water seals.
* Make water seals. In 6.5mm steel plate, drill a hole slightly larger than the shaft diameter (about
4.53cm). Make one for each side. Weld or bolt to the inside of the turbine housing. The shaft must pass

through the seals without touching them. Some water will still come through the housing but not
enough to interfere with efficiency.
* Make the foundation to which the bearings will be attached of hardwood pilings or concrete.
* Move the turbine, with bearings attached, to the proper nozzle/turbine placement and attach the
bearings to the foundation with bolts. The bearings will be on the outside of the turbine housing (see
Figure 21). (Note: The drive pulley is omitted from the Figure for clarity.)

Figure 22 shows a possible turbine installation for high head applications. A water shut-off valve allows
control of the flow of water. Never shut off the water flow suddenly as a rupture in the penstock is
certain to occur. If maintenance on the turbine is necessary, reduce the flow gradually until the water
stops.

VI. MAINTENANCE
The Michell (Banki) turbine is relatively maintenance-free. The only wearable parts are the bearings
which may have to be replaced from time to time. An unbalanced turbine or a turbine that is not
mounted exactly will wear the bearings very quickly.
A chicken wire screen (1.5cm x 1.5cm weave) located behind the control gate will help to keep branches
and rocks from entering the turbine housing. It may be necessary to clean the screen from time to time.
An alternative to chicken wire is the use of thin steel rods spaced so that a rake can be used to remove
any leaves or sticks.
VII. ELECTRICAL GENERATION
It is beyond the scope of this article to go into electrical generation using the Michell (Banki) turbine.
Depending on the generator and accessories you choose, the turbine can provide enough rpm for direct
current (DC) or alternating current (AC).
For information on the type of generator to purchase, contact manufacturers directly. A list of
companies is provided here. The manufacturer often will be able to recommend an appropriate
generator, if supplied with enough information upon which to make a recommendation. Be prepared to
supply the following details:

* AC or DC operation (include voltage desired).


* Long range use of electrical energy (future consumption and addition of electric devices).
* Climatic condition under which generator will be used (i.e., tropical, temperate, arid, etc.).
* Power available at water site calculated at lowest flow and maximum flow rates.
* Power available to the generator in watts or horsepower (conservative figure would be half of power
at water site).
* Revolutions per minute (rpm) of turbine without pulleys and belt.
* Intended or present consumption of electrical energy in watts if possible (include frequency of
electrical use).
VIII. DICTIONARY OF TERMS
AC (Alternating Current)--Electrical energy that reverses its direction at regular intervals. These intervals
are called cycles.
BEARING--Any part of a machine in or on which another part revolves, slides, etc.
DIA (Diameter)--A straight line passing completely through the center of a circle.
DC (Direct Current)--Electrical current that flows in one direction without deviation or interruption.
GROSS POWER--Power available before machine inefficiencies are subtracted.
HEAD--The height of a body of water, considered as causing pressure.
ID (Inside Diameter)--The inside diameter of pipe, tubing, etc.
NET HEAD--Height of a body of water minus the energy losses caused by the friction of a pipe or water
channel.
OD (Outside Diameter)--The outside dimension of pipe, tubing, etc.
PENSTOCK--A conduit or pipe that carries water to a water wheel or turbine.
ROLLED EARTH--Soil that is pressed together tightly by rolling a steel or heavy wood cylinder over it.
RPM (Revolutions Per Minute)--The number of times something turns or revolves in one minute.
TAILRACE (Tailwater)--The discharge channel that leads away from a waterwheel or turbine.
TURBINE--Any of various machines that has a rotor that is driven by the pressure of such moving fluids
as steam, water, hot gases, or air. It is usually made with a series of curved blades on a central rotating
spindle.
WEIR--A dam in a stream or river that raises the water level.
X. CONVERSION TABLES
UNITS OF LENGTH
1 Mile = 1760 Yards = 5280 Feet
1 Kilometer = 1000 Meters = 0.6214 Mile
1 Mile = 1.607 Kilometers
1 Foot = 0.3048 Meter
1 Meter = 3.2808 Feet = 39.37 Inches
1 Inch = 2.54 Centimeters
1 Centimeter = 0.3937 Inches
UNITS OF AREA
1 Square Mile = 640 Acres = 2.5899 Square Kilometers
1 Square Kilometer = 1,000,000 Square Meters = 0.3861 Square Mile
1 Acre = 43,560 Square Feet
1 Square Foot = 144 Square Inches = 0.0929 Square Meter
1 Square Inch = 6.452 Square Centimeters
1 Square Meter = 10.764 Square Feet
1 Square Centimeter = 0.155 Square Inch
UNITS OF VOLUME
1.0 Cubic Foot = 1728 Cubic Inches = 7.48 US Gallons
1.0 British Imperial Gallon = 1.2 US Gallons
1.0 Cubic Meter = 35.314 Cubic Feet = 264.2 US Gallons
1.0 Liter = 1000 Cubic Centimeters = 0.2642 US Gallons

UNITS OF WEIGHT
1.0 Metric Ton = 1000 Kilograms = 2204.6 Pounds
1.0 Kilogram = 1000 Grams = 2.2046 Pounds
1.0 Short Ton = 2000 Pounds
UNITS OF PRESSURE
1.0 Pound per square inch = 144 Pound per square foot
1.0 Pound per square inch = 27.7 Inches of water*
1.0 Pound per square inch = 2.31 Feet of water*
1.0 Pound per square inch = 2.042 Inches of mercury*
1.0 Atmosphere = 14.7 Pounds per square inch (PSI)
1.0 Atmosphere = 33.95 Feet of water*
1.0 Foot of water = 0.433 PSI = 62.355 Pounds per square foot
1.0 Kilogram per square centimeter = 14.223 Pounds per square inch
1.0 Pound per square inch = 0.0703 Kilogram per square centimeter
UNITS OF POWER
1.0 Horsepower (English) = 746 Watt = 0.746 Kilowatt (KW)
1.0 Horsepower (English) = 550 Foot pounds per second
1.0 Horsepower (English) = 33,000 Foot pounds per minute
1.0 Kilowatt (KW) = 1000 watt = 1.34 Horsepower (HP) English
1.0 Horsepower (English) = 1.0139 Metric horsepower (cheval-vapeur)
1.0 Metric horsepower = 75 Meter X Kilogram/Second
1.0 Metric horsepower = 0.736 Kilowatt = 736 Watt (*)
At 62 degrees Fahrenheit (16.6 degrees Celsius).
APPENDIX I SITE ANALYSIS
This Appendix provides a guide to making the necessary calculations for a detailed site analysis.
Data Sheet Measuring, Gross Head Measuring, Flow Measuring, Head Losses
DATA SHEET
1. Minimum flow of water available in cubic feet per second (or cubic meters per second). _____
2. Maximum flow of water available in cubic feet _____ per second (or cubic meters per second).
3. Head or fall of water in feet (or meters). _____
4. Length of pipe line in feet (or meters) needed to get the required head. _____
5. Describe water condition (clear, muddy, sandy, acid). _____
6. Describe soil condition (see Table 2). _____
7. Minimum tailwater elevation in feet (or meters). _____
8. Approximate area of pond above dam in acres (or square kilometers). _____
9. Approximate depth of the pond in feet (or meters). _____
10. Distance from power plant to where electricity will be used in feet (or meters). _____
11. Approximate distance from dam to power plant. _____
12. Minimum air temperature. _____
13. Maximum air temperature. _____
14. Estimate power to be used. _____
ATTACH SITE SKETCH WITH ELEVATIONS, OR TOPOGRAPHICAL MAP WITH SITE SKETCHED IN.
The following questions cover information which, although not necessary in starting to plan a water
power site, will usually be needed later. If it can possibly be given early in the project, this will save time
later.
1. Give the type, power, and speed of the machinery to be driven and indicate whether direct, belt, or
gear drive is desired or acceptable.
2. For electric current, indicate whether direct current is acceptable or alternating current is required.
Give the desired voltage, number of phases and frequency.
3. Say whether manual flow regulation can be used (with DC and very small AC plants) or if regulation by
an automatic governor is needed.

MEASURING GROSS HEAD


Method No. 11. Equipment

a. Surveyor's leveling instrument--consists of a spirit level fastened parallel to a telescopic sight. b. Scale-use wooden board approximately 12 ft in length. 2. Procedure a. Surveyor's level on a tripod is placed
downstream from the power reservoir dam on which the headwater level is marked. b. After taking a
reading, the level is turned 180[degrees] in a horizontal circle. The scale is placed downstream from it at
a suitable distance and a second reading is taken. This process is repeated until the tailwater level is
reached.
MEASURING HEAD WITH SURVEYOR'S LEVEL

Method No. 2 This method is fully reliable, but is more tedious than Method No. 1 and need only be
used when a surveyor's level is not available. 1. Equipment

a. Scale b. Board and wooden plug c. Ordinary carpenter's level 2. Procedure a. Place board horizontally
at headwater level and place level on top of it for accurate leveling. At the downstream end of the
horizontal board, the distance to a wooden peg set into the ground is measured with a scale. b. The
process is repeated step by step until the tailwater level is reached.
MEASURING HEAD WITH CARPENTER'S LEVEL

MEASURING FLOW
Flow measurements should take place at the season of lowest flow in order to guarantee full power at
all times. Investigate the stream's flow history to determine the level of flow at both maximum and
minimum. Often planners overlook the fact that the flow in one stream may be reduced below the
minimum level required.
Other streams or sources of power would then offer a better solution. Method No. 1 For streams with a
capacity of less than one cubic foot per second, build a temporary dam in the stream, or use a
"swimming hole" created by a natural dam. Channel the water into a pipe and catch it in a bucket of
known capacity.
Determine the stream flow by measuring the time it takes to fill the bucket. Stream flow (cubic ft/sec) =
Volume of bucket (cubic ft) Filling time (seconds) Method No. 2 For streams with a capacity of more
than 1 cu ft per second, the weir method can be used.
The weir is made from boards, logs, or scrap lumber. Cut a rectangular opening in the center. Seal the
seams of the boards and the sides built into the banks with clay or sod to prevent leakage. Saw the
edges of the opening on a slant to produce sharp edges on the upstream side. A small pond is formed
upstream from the weir. When there is no leakage and all water is flowing through the weir opening, (1)
place a board across the stream and (2) place another narrow board at right angles to the first, as shown
below. Use a carpenter's level to be sure the second board is level.

FIGURE A

Measure the depth of the water above the bottom edge of the weir with the help of a stick on which a
scale has been marked. Determine the flow from Table 1 on page 56.
FIGURE B

Table 1
FLOW VALUE
Weir Width
(Feet/Second)
Overflow Height3 feet4 feet5 feet6 feet7 feet8 feet9 feet
1.0 inch
0.24 0.32 0.40 0.48 0.56 0.64 0.72
2.0 inches
0.67 0.89 1.06 1.34 1.56 1.80 2.00
4.0 inches
1.90 2.50 3.20 3.80 4.50 5.00 5.70
6.0 inches
3.50 4.70 5.90 7.00 8.20 9.40 10.50
8.0 inches
5.40 7.30 9.00 10.90 12.40 14.60 16.20
10.0 inches
7.60 10.00 12.70 15.20 17.70 20.00 22.80
12.0 inches
10.00 13.30 16.70 20.00 23.30 26.60 30.00

Method No. 3
The float method is used for larger streams. Although it is not as accurate as the previous two methods,
it is adequate for practical purposes. Choose a point in the stream where the bed is smooth and the
cross section is fairly uniform for a length of at least 30 ft. Measure water velocity by throwing pieces of
wood into the water and measuring the time of travel between two fixed points, 30 ft or more apart.
Erect posts on each bank at these points.
Connect the two upstream posts by a level wire rope (use a carpenter's level). Follow the same
procedure with the downstream posts. Divide the stream into equal sections along the wires and
measure the water depth for each section. In this way, the cross-sectional area of the stream is
determined. use the following formula to calculate the flow:
FIGURE C

MEASURING HEAD LOSSES


"Net Power" is a function of the "Net Head." The "Net Head" is the "Gross Head" less the "Head Losses."
The illustration below shows a typical small water power installation. The head losses are the openchannel losses plus the friction loss from flow through the penstock.
FIGURE D

FIGURE E
Open Channel Head Losses The headrace and the tailrace in the illustration above are open channels for
transporting water at low velocities. The walls of channels made of timber, masonry, concrete, or rock,
should be perpendicular. Design them so that the water level height is one-half of the width. Earth walls
should be built at a 45 [degrees] angle.

Design them so that the water level height is one-half of the channel width at the bottom. At the water
level the width is twice that of the bottom. The head loss in open channels is given in the nomograph.
The friction effect of the material of construction is called "N." Various values of "N" and the maximum
water velocity, below which the walls of a channel will not erode are given.
TABLE II
Maximum Allowable
Material of Channel Wall
Water
VelocityValue of "n"
(feet/second)
Fine grained sand
0.6
0.030
Course sand
1.2
0.030
Small stones
2.4
0.030
Coarse stones
4.0
0.030
Rock (Smooth)
25.0
0.033
Rock (Jagged)
25.0
0.045
Concrete with sandy water
10.0
0.016
Concrete with clean water
20.0
0.016
Sandy loam, 40% clay
1.8
0.030
Loamy soil, 65% clay
3.0
0.030
Clay loam, 85% clay
4.8
0.030
Soil loam, 95% clay
6.2
0.030
100% clay
7.3
0.030
Wood
0.015
Earth bottom with rubble sides
0.033
The hydraulic radius is equal to a quarter of the channel width, except for earth-walled channels where
it is 0.31 times the width at the bottom. To use the nomograph, a straight line is drawn from the value
of "n" through the flow velocity to the reference line. The point on the reference line is connected to
the hydraulic radius and this line is extended to the head-loss scale which also determines the required
slope of the channel.
Using a Nomograph
After carefully determining the water power site capabilities in terms of water flow and head, the
nomograph
is
used
to
determine:
* The width/depth of the channel needed to bring the water to the spot/location of the water turbine.
* The amount of head lost in doing this.
FIGURE F

To use the graph, draw a straight line from the value of "n" through the flow velocity through the
reference line tending to the hydraulic radius scale.
The hydraulic radius is one-quarter (0.25) or (0.31) the width of the channel that needs to be built. In
the case where "n" is 0.030, for example, and water flow is 1.5 cubic feet/second, the hydraulic radius is
0.5 feet hr 6 inches. If you are building a timber, concrete, masonry, or rock channel, the total width of
the channel would be 6 inches times 0.25, or 2 feet with a depth of at least 1 foot.
If the channel is made of earth, the bottom width of the channel would be 6 times 0.31, or 19.5 inches,
with a depth of at least 9.75 inches and top width of 39 inches.
Suppose, however, that water flow is 4 cubic feet/second. Using the graph, the optimum hydraulic
radius would be approximately 2 feet--or for a wood channel, a width of 8 feet. Building a wood channel
of this dimension would be prohibitively expensive.
FIGURE G

However, a smaller channel can be built by sacrificing some water head. For example, you could build a
channel with a hydraulic radius of 0.5 feet or 6 inches. To determine the amount of head that will be
lost, draw a straight line from the value of "n" through the flow velocity of 4 [feet 3]/second to the
reference line.
Now draw a straight line from the hydraulic radius scale of 0.5 feet through the point on the reference
line extending this to the head-loss scale which will determine the slope of the channel. In this case
about 10 feet of head will be lost per thousand feet of channel. If the channel is 100 feet long, the loss
would only be 1.0 feet--if 50 feet long, 0.5 feet, and so forth.
Pipe Head Loss and Penstock Intake
The trashrack consists of a number of vertical bars welded to an angle iron at the top and a bar at the
bottom (see Figure below). The vertical bars must be spaced so that the teeth of a rake can penetrate
the rack for removing leaves, grass, and trash which might clog up the intake. Such a trashrack can easily
be manufactured in the field or in a small welding shop. Downstream from the trashrack, a slot is
provided in the concrete into which a timber gate can be inserted for shutting off the flow of water to
the turbine. (See shut-off caution on page 31.)
FIGURE H

The penstock can be constructed from commercial pipe. The pipe must be large enough to keep the
head loss small. The required pipe size is determined from the nomograph. A straight line drawn
through the water velocity and flow rate scales gives the required pipe size and pipe head loss. Head
loss is given for a 100-foot pipe length.
For longer or shorter penstocks, the actual head loss is the head loss from the chart multiplied by the
actual length divided by 100. If commercial pipe is too expensive, it is possible to make pipe from native
material; for example, concrete and ceramic pipe, or hollowed logs. The choice of pipe material and the
method of making the pipe depend on the cost and availability of labour and the availability of material.
FIGURE I

APPENDIX II
SMALL DAM CONSTRUCTION
Introduction to: Earth Dams, Crib Dams, Concrete and Masonry Dams
This appendix is not designed to be exhaustive; it is meant to provide background and perspective for
thinking about and planning dam efforts. While dam construction projects can range from the simple to
the complex, it is always best to consult an expert, or even several; for example, engineers for their
construction savvy and an environmentalist or concerned agriculturalist for a view of the impact of
damming.
EARTH DAMS
An earth dam may be desirable where concrete is expensive and timber scarce. It must be provided with
a separate spillway of sufficient size to carry off excess water because water can never be allowed to
flow over the crest of an earth dam. Still water is held satisfactorily by earth but moving water is not.
The earth will be worn away and the dam destroyed.
The spillway must be lined with boards or concrete to prevent seepage and erosion. The crest of the
dam may be just wide enough for a footpath or may be wide enough for a roadway, with a bridge
placed across the spillway.
FIGURE J

The big problem in earth-dam construction is in places where the dam rests on solid rock. It is hard to
keep the water from seeping between the dam and the earth and finally undermining the dam.
One way of preventing seepage is to blast and clean out a series of ditches, or keys, in the rock, with
each ditch about a foot deep and two feet wide extending under the length of the dam. Each ditch
should be filled with three or four inches of wet clay compacted by stamping it. More layers of wet clay
can then be added and the compacting process repeated each time until the clay is several inches higher
than bedrock.
The upstream half of the dam should be of clay or heavy clay soil, which compacts well and is
impervious to water. The downstream side should consist of lighter and more porous soil which drains
quickly and thus makes the dam more stable than if it were made entirely of clay.
EARTH-FILL DAM

CRIB DAMS
The crib dam is very economical where lumber is easily available: it requires only rough tree trunks, cut
planking, and stones. Four- to six-inch tree trunks are placed 2-3 feet apart and spiked to others placed
across them at right angles. Stones fill the spaces between timbers.
The upstream side (face) of the dam, and sometimes the downstream side, is covered with planks. The
face is sealed with clay to prevent leakage. Downstream planks are used as an apron to guide the water
that overflows the dam back into the stream bed.
The dam itself serves as a spillway in this case. The water coming over the apron falls rapidly. Prevent
erosion by lining the bed below with stones. The apron consists of a series of steps for slowing the water
gradually.
FIGURE K

FIGURE L
Crib dams must be embedded well into the embankments and packed with impervious material such as
clay or heavy earth and stones in order to anchor them and to prevent leakage. At the heel, as well as at
the toe of crib dams, longitudinal rows of planks are driven into the stream bed. These are priming
planks which prevent water from seeping under the dam. They also anchor the dam.

If the dam rests on rock, priming planks cannot and need not be driven; but where the dam does not
rest on rock they make it more stable and watertight. These priming planks should be driven as deep as
possible and then spiked to the timber of the crib dam.
The lower ends of the priming planks are pointed as shown in the Figure on page 69 and must be placed
one after the other as shown. Thus each successive plank is forced, by the act of driving it, closer against
the preceding plank, resulting in a solid wall. Any rough lumber may be used.

Chestnut and oak are considered to be the best material. The lumber must be free from sap, and its size
should be approximately 2" X 6".
In order to drive the priming planks, considerable force may be required. A simple pile driver will serve
the purpose. The Figure below shows an excellent example of a pile driver.

CONCRETE AND MASONRY DAMS


Concrete and masonry dams more than 12 feet high should not be built without the advice of an
engineer with experience in this field. Dams require knowledge of the soil condition and bearing
capacity as well as of the structure itself.
A stone dam can also serve as a spillway. It can be up to 10 feet in height. It is made of rough stones.

The layers should be bound by concrete. The dam must be built down to a solid and permanent footing
to prevent leakage and shifting.
The base of the dam should have the same dimensions as its height to give it stability.
Small concrete dams should have a base with a thickness 50 percent greater than height. The apron is
designed to turn the flow slightly upwards to dissipate the energy of the water and protect the
downstream bed from erosion.
SMALL CONCRETE DAM

APPENDIX III
DECISION MAKING WORKSHEET
If you are using this as a guide for using the Michell (Banki) Turbine in a development effort, collect as
much information as possible and if you need assistance with the project, write VITA. A report on your
experiences and the uses of this Manual will help VITA both improve the book and aid other similar
efforts. Volunteers in Technical Assistance 1600 Wilson Boulevard, Suite 500 Arlington, Virginia 22209,
USA
CURRENT USE AND AVAILABILITY * Describe current agricultural and domestic practices which rely on
water. What are the sources of water and how are they used?
* What water power sources are available? Are they small but fast-flowing? Large but slow-flowing?
Other characteristics?
* What is water used for traditionally?
* Is water harnessed to provide power for any purpose? If so, what and with what positive or negative
results?
* Are there dams already built in the area? If so, what have been the effects of the damming? Note

particularly any evidence of sediment carried by the water--too much sediment can create a swamp.
* If water resources are not now harnessed, what seem to be the limiting factors? Does cost seem
prohibitive? Does the lack of knowledge of water power potential limit its use?
NEEDS AND RESOURCES
* Based on current agricultural and domestic practices, what seem to be the areas of greatest need? Is
power needed to run simple machines such as grinders, saws, pumps?
* Given available water power sources, which ones seem to be available and most useful? For example,
one stream which runs quickly year around and is located near the center of agricultural activity may be
the only feasible source to tap for power.
* Define water power sites in terms of their inherent potential for power generation.
* Are materials for constructing water power technologies available locally? Are local skills sufficient?
Some water power applications demand a rather high degree of construction skill.
* What kinds of skills are available locally to assist with construction and maintenance? How much skill
is necessary for construction and maintenance? Do you need to train people? Can you meet the
following needs?
* Some aspects of the Michell turbine require someone with experience in metalworking and/or
welding.
* Estimated labour time for full-time workers is:
* 40 hours skilled labour
* 40 hours unskilled labour
* 8 hours welding
* Do a cost estimate of the labour, parts, and materials needed.
* How will the project be funded?
* What is your schedule? Are you aware of holidays and planting or harvesting seasons which may affect
timing?
* How will you arrange to spread information on and promote use of the technology?
IDENTIFY POTENTIAL
* Is more than one water power technology applicable? Remember to look at all the costs. While one
technology appears to be much more expensive in the beginning, it could work out to be less expensive
after all costs are weighed.
* Are there choices to be made between a waterwheel and a windmill, for example, to provide power
for grinding grain? Again weigh all the costs: economics of tools and labour, operation and maintenance,
social and cultural dilemmas.
* Are there local skilled resources to introduce water power technology? Dam building and turbine
construction should be considered carefully before beginning work. Besides the higher degree of skill
required in turbine manufacture (as opposed to waterwheel construction), these water power
installations tend to be more expensive.
* Where the need is sufficient and resources are available, consider a manufactured turbine and a group
effort to build the dam and install the turbine. * Is there a possibility of providing a basis for small
business enterprise?
FINAL DECISION * How was the final decision reached to go ahead--or not go ahead--with this
technology? Why?
APPENDIX IV
RECORD KEEPING WORKSHEET CONSTRUCTION Photographs of the construction process, as well as the
finished result, are helpful. They add interest and detail that might be overlooked in the narrative. A
report on the construction process should include much very specific information. This kind of detail can
often be monitored most easily in charts (such as the one below).
CONSTRUCTION
Labor Account Hours Worked Name Job M T W T F S S Total Rate? Pay? 1 2 3 4 5 Totals Materials
Account Item Cost Per Item # Items Total Costs 1 2 3 4 5 Total Costs Some other things to record
include: * Specification of materials used in construction. * Adaptations or changes made in design to fit

local conditions. * Equipment costs. * Time spent in construction--include volunteer time as well as paid
labour; full- or part-time. * Problems--labour shortage, work stoppage, training difficulties, materials
shortage, terrain, transport.
OPERATION
Keep log of operations for at least the first six weeks, then periodically for several days every few
months. This log will vary with the technology, but should include full requirements, outputs, duration
of operation, training of operators, etc. Include special problems that may come up--a damper that
won't close, gear that won't catch, procedures that don't seem to make sense to workers, etc.
MAINTENANCE
Maintenance records enable keeping track of where breakdowns occur most frequently and may
suggest areas for improvement or strengthening weakness in the design. Furthermore, these records
will give a good idea of how well the project is working out by accurately recording how much of the
time it is working and how often it breaks down. Routine maintenance records should be kept for a
minimum of six months to one year after the project goes into operation.
MAINTENANCE Labor Account Also down time Name Hours & Date Repair Done Rate? Pay? 1 2 3 4 5
Totals (by week or month) Materials Account Item Cost Reason Replaced Date Comments 1 2 3 4 5
Totals (by week or month) SPECIAL COSTS This category includes damage caused by weather, natural
disasters, vandalism, etc. Pattern the records after the routine maintenance records. Describe for each
separate incident: * Cause and extent of damage. * Labor costs of repair (like maintenance account). *
Material costs of repair (like maintenance account). * Measures taken to prevent recurrence.
<FIGURE M>

1-kW RIVER GENERATOR


by MATHEW G. BOISSEVAIN
The plan presented here is a detailed description of a 1-kilowatt (1-kW) generator unit, which was
prepared in 1971. The plan is scheduled to be revised and updated in the near future in order to
incorporate additional data. At the time, this plan was prepared, the generator had not been built on the
scale shown here. Therefore, until such time as testing results can be integrated into the plan, VITA offers
this material as an idea paper.
By way of background, the designer of the 1-kW river generator made the following assumption in his
calculations:

80% efficiency for each of the three "V"-belt speed-up stages so that enough power is available to operate
the unit at 4.7 ft/sec water velocity. It will certainly operate at 6.0 ft/sec.

WHAT IT IS
The river generator uses the flow of river water to produce 1-kW of electrical power. It has four 5-ft
diameter propellers attached to a log float. The float is anchored to the river bottom.
From actual trials with 40" propellers, it is calculated that water flowing at 4.7 to 6.0 ft/sec will turn the
5-ft propellers with enough power to generate 1000 watts of electricity.
The propellers are connected to the generator by a series of "V"-belt pulley drives that speed up the slow
66-revolutions per minute (rpm) of the propellers to the fast 3600-rpm of the generator.
A variable speed pulley at the generator and a built-in volt meter will enable the user to adjust the
generator rpm and voltage for varying river flow and generator load conditions. Most of the pulleys,
shafts, and "V"-belts are identical. This simplifies building and reduces spare parts needed.
Due to slow speeds (except at the last stage) the unit should last a long time, if built as instructed. It does
not need elabourate dams, river falls, or pipes, as do Most hydroelectric plants. You can build as many of
these generators as needed, spaced about 100 feet apart downstream. By comparison, a paddle wheel,
under identical conditions, would need paddles at least as large as the entire back frame of the river
generator (see drawing, rear view) to provide the same amount of power. Also, it would only turn at about
5-7 rpm and would need more speed-up stages of "V"-belts.
The 1-kW river generator will produce about 720-kW hours per month, which should be enough energy
to run a simple household.
Some effort must be made to conserve energy, and to spread energy use as evenly through the day as
possible.

WHAT YOU NEED TO BUILD ONE


This design requires access to a river that runs with a speed of from 4.7 to 6.0-ft/sec year round with a
depth of at least 6-ft over a 21-ft width. This principle will also work on a smaller scale with a
corresponding reduction in power output. If you have access to a waterfall or higher water speeds, you
could make a 1-kW or bigger unit with smaller (but stronger) propellers, turning at higher speeds, and
with fewer pulleys and belts. These higher speeds can also be obtained by building dams, etc.
Purchased parts (see parts list) total US$612.81, based on 1971 list prices in the United States. To this,
you must add tax and shipping and the cost of all lumber and logs used. To reduce cost, you may be able
to find your own parts (airplane propellers, large fans, washing machine pulleys, etc.). Because prices
may have changed drastically since 1971, be sure it is economically feasible before you begin
construction.

Tools
* Wood saw
* Hammer and/or hatchet
* 1" wood drill
* 1/2" wood drill
* 3/16" metal drill (for nail holes in item 21)
* Allen wrenches (for 1/4" set screws in pulleys)
* Metal file
* Pliers to cut and twist 1/8" wire
* Wrench for 1/4" bolts in propellers
* .669" diameter drill for increasing hole in cast iron pulley (item 13) from .625" diameter to .669/.673"
diameter.

Lumber
* 14 straight logs, 5"-8" diameter X 21' (Bamboo may also be used instead of logs. Use wire around joints
to make strong, durable structure.)
* Planks, 2" X 4"; as shown on page 9.

WHERE TO BUILD

Find a place in your river, close to home, that has a water speed of 4.7 to 6.0-ft/sec. This must be
measured exactly. (At 3-ft/sec water speed, you will get only about one-fourth of the power, or 250
watts.) To measure the speed, measure off 50-feet along the river bank and mark with sticks. Then, count
the exact time required for the water to carry a log or branch between the sticks. The time must be 8.3 to
10.6 seconds. The place to build your river generator is upstream from the location chosen, so it can be
launched like a large boat and floated into place. After the frame has been built, attach the rope (item 20)
to the pointed end and anchor the float in place, using a boat anchor, a large rock, or a 2-inch steel pipe
driven into the river bottom.

HOW TO BUILD
Float. Before spending money on parts, it may be wise to build the log frame and test it out to make sure
it holds together under all river-flood conditions. The author has successfully built smaller frames than
the one shown. Unforeseen problems may arise with a larger one.
Build the frame as shown in the following drawings. Be sure to reinforce all the joints with metal plates
(item 21). (Tin cans could possibly be used here. Remove ends, flatten, then dip in paint to prevent
rusting.) The wire (item 22) is used to keep branches, etc., out of the propellers. Also, use wire braces at
the back section (see rear view), and on all joints when building the frame with bamboo instead of logs.
Drive System. Next, make all the wooden parts shown on the "side view" drawing. Items 18 and 19
should be made of hardwood. Soak it in water after drilling the hole. Then open up the hole as needed
until the 1-inch shaft turns smoothly. Do the same for item 3 (pivoting frames--to provide V-belt tension),
but soak it in oil around the hole where the shaft goes by filling oil holes after shaft is assembled. Enlarge
if necessary to provide smooth running shaft.
Attach item 9 to the vertical members on the float, with the shaft in place. Make sure the shaft is square
with the frame as shown in the drawing. It is important to end up with the dimensions as shown in the
"rear view" drawing, because the "V"-belts are not adjustable in length.
Shafts. Attach the pulleys to all the shafts, as shown. File or drill 1/8-inch into the shaft under all the set
screws. This prevents the pulleys from turning on the shaft. The propellers attach with a split bushing and
will grip the shaft tightly when the bushing is bolted tightly against the propeller hub.
Belts. Install all propeller belts and weight down each item 3 on both sides of center with rocks, as
shown. This will provide a constant, even tension on the belts and reduce belt wear. Allow the belts to set
in the exact position of item 12 before bracing it as shown on the "rear view" drawing.
Repeat above with the tall item 12 at the center of the float. After the unit is back in the water and
operating under load, add rocks as needed to prevent belt slippage. Don't make the belts too tight as this
increases wear.
Adjust the variable speed pulley on the generator to get the 120-volts needed. Be careful to ground the
generator housing and ground wire in the water, and to keep people away who don't know the dangers of
120-volts of electricity and wet feet in water, etc.
Insulate all exposed wire. Follow wiring instructions that come with the generator, and run the cable
down the anchor rope, then along the river bottom (weight with rocks) to the shore. Propeller Blade Tip
Angle Adjustment. At 4.7-ft/sec water velocity, the propeller must turn at 60 to 70-rpm. Twist the
propeller tip until the blade is angled as shown below. All blades must have a 9 [degrees] angle.

FIGURE 1

SUGGESTED SUPPLIERS
Grainger W. W. Grainer, Inc., 519 Potrero, San Francisco, California U.S.A. [Phone: (415) 861-48411]
Browning Browning Mfg. Division, Emerson Electric Co., Maysville, Kentucky 41056 U.S.A.
Ryerson Ryerson Steel, Box 8427, Emeryville, California 94608 U.S.A. (also, U.S. Steel)
Durkee Atwood Durkee Atwood Co., Minneapolis, Minnesota 55413 U.S.A. [Phone: (612) 332-0441]
Sears, Roebuck & Co. Los Angeles, California 90054 U.S.A. (NOTE: Drill item 13 from .625" diameter
to .669" diameter.)

PARTS LIST
ItemNumber Quantity
Needed

Description

Stock
Number

Cost
Each

Total
Cost

Where
Buy

60" exhaust fan, 1" bore

3 CO 32

41.35

165.40

Grainger

14

14" pulley, 1" bore

3 X 944

3.82

53.48

Grainger

14

4" pulley, 1" bore

AS-40

2.53

35.42

Browning

14

"A" section "V"-belt

A 158

6.43

90.02

Durkee
Atwood

Alternator, 1.2-kW, 115- F32


KF 119.00
Volt
32054 N

119.00

Sears

1" X 18" long shaft

61.00

Ryerson

3" door hinge

1.20

1.20

(local)

250 ft

Underground cable,
gauge, 5/8" bore

26.75

26.75

Type 303 SS 61.00


12 1 W 676

Grainger

13

Variable pitch pulley

3 X 276

2.77

5.54

Grainger

15

"A" section "V"-belt

A-75

5.00

10.00

Durkee
Atwood

17

36 (14#)

4" X 8" X .125" thick

6061-T6
alum

8.00

Ryerson

18

12

1" flat washer

plated steel

1.00

(local)

20

50 ft

1" diameter rope

25.00

(local)

21

20 lb

4" long nails, plated wire

6.00

(local)

22
300 ft
1/8" soft galvanized steel
Total cost of all purchased parts except wood $612.81
Add tax and shipping

5.00

Ryerson

to

FIGURE

FIGURE 3

Click

image

for

larger

view

FIGURE

Click

image

for

larger

view

Mathew G. Boissevain is a design engineer at a major U.S. corporation. During his many years as a
VITA Volunteer, he has developed several types of water wheels for powering water pumps and has
worked for almost 20 years designing machines used in automated processes--for example, artificial
kidney equipment, circular weaving loom, various mail-sorting devices, food-processing machines.
Please send testing results, comments, suggestions, and requests for further information to:
VITA 1600 Wilson Boulevard, Suite 500 Arlington, Virginia 22209 USA Tel: 703/276-1800 * Fax:
703/243-1865 Internet: pr-info@vita.org 0-86619-079-1
VITA Technical Bulletins offer do-it-yourself technology information on a wide variety of subjects.
The Bulletins are idea generators intended not so much to provide a definitive answer as to guide the
user's thinking and planning. Premises are sound and testing results are provided, if available.
Evaluations and comments based on each user's experience are requested. Results are incorporated
into subsequent editions, thus providing additional guidelines for adaptation and use in a greater
variety of conditions.

A Design Manual for Water Wheels


With details for applications to pumping water for village use and driving small machinery
WILLIAM G. OVENS

PART ONE: THE WATER WHEEL


I. INTRODUCTION
Supplying power to many remote locations in the world from central generators using customary
distribution methods is either economically unfeasible or will be many years in coming. Power, where
desirable, will therefore need to be generated locally. Various commercial machinery is marketed, but the
required capital expenditure or maintenance/running cost is beyond the capability of many potential
users.
Some effort has been expended at the Papua New Guinea University of Technology to devise low cost
means of generating modest amounts of power in remote locations. This paper reports on one such project
involving the development of low cost machinery to provide mechanical power. Regardless of the final
use to which the power is put the natural sources of energy which can be utilized are fairly readily
categorized. Among them:
1. Falling water
2. Animals
3. Sun
4. Wind
5. Fossil fuels
6. Nuclear fuels
7. Organic waste
Sun, wind and water are free and renewable in the sense that by using them we do not alter their future
usefulness. From continually operating cost considerations, a choice from among these is attractive. From
capital cost consideration hydro-power may be very unattractive. Sun and wind have obvious natural
limitations based upon local weather conditions.
Furthermore, for technological and economic reasons, solar power use is presently limited to applications
utilizing the energy directly as part of a heat cycle. Animals require specialized care and continuous food
sources. Conversion of organic waste to useable energy is being experimented with, with varying success,
in several parts of the world.
Whatever the form of the naturally occurring energy, it may be transformed, if necessary, into useable
power in a wide variety of ways. The choice of method depends upon a complex interaction of too many
considerations to enumerate fully here, but among them are:
1. the use to which the power will be put;
2. the form in which it will be utilized. This generally, but not exclusively, falls into the broad categories
of mechanical and electrical;
3. the economic and natural resources available;
4. availability of suitable maintenance facilities;
5. whether the machinery must be portable or not.

II. FORMULATION OF THE PROBLEM


In the absence of a specific request from government or any outside body, the decision was taken based
primarily on the obvious abundance of available water power to investigate broadly the design
possibilities for low cost machinery to produce small amounts of mechanical power. One immediately
obvious potential application is the generation of electric power, but for reasons mentioned under "Other
Applications" in Part Two this has not been pursued.
However, in many places, villages are located at some distance from the traditional source of drinking
water. The principal intended use for the power generated by the machine discussed in this manual has
been the pumping of potable water for distribution to a village. The project, thus, has included
construction of a simple pump attachment also. Several other potential uses are discussed later.
Limits on the scope of the project were decided based upon numerous considerations:
1. Minimum of capital expenditure indicated a device which could be constructed locally of inexpensive
materials with no specialized, expensive components or machinery required.

2. Local construction suggested the desirability of design details requiring only simple construction
techniques.
3. Since the installation was likely to be remote (indicating a probable shortage of local skilled
tradesmen) maintenance, if any, would have to be minimal and simple.
4. The device should be such that repair, if any, could be carried out on-site with parts and necessary tools
light enough to be carried easily to the site.
5. The usual considerations of safety must apply with the knowledge that the village children could
not/would not be kept away from the device.
I decided to concentrate on investigating the feasibility of using the water wheel, it being the device
which seemed most likely to optimize the criteria set out above. There are other types of machines
suitable for creating mechanical power from hydro sources, but none, known to me, can be constructed
with such simple techniques requiring so low a level of trade skills as the wooden water wheel.
Water wheels are in use in various parts of the world now. Many have been constructed on an ad hoc
basis and vary in complexity, efficiency and ingenuity of design and construction. The basic device is so
simple that a workable wheel can be constructed by almost anyone who has the desire to try. However,
the subtleties of design which separate adequate from inadequate models may escape those without
sufficient technical training.
The number of projects abandoned after a relatively short life attests to the fact that designers/builders
often have more pluck than skill. It seems desirable to attack the problem in a systematic fashion with an
objective of establishing a design manual for the selection of proper sizes required to meet a specific need
and to set out design features based on sound engineering principles. I offer the following as an attempt to
meet that objective.
The wheel consists of buckets-to hold the water-fixed in a frame and arranged so that buckets and frame
together rotate about a centre axis which is oriented perpendicular to the inlet water flow. Traditional
designs employ the undershot, overshot or breast configurations. In the undershot wheel, the inlet water
flows tangent to the bottom edge of the wheel. In the overshot wheel, the water is brought in tangent to
the top edge of the wheel, partially or fully filling the bucket.
It is carried in the buckets until dumped out somewhat before reaching the lowest point on the wheel. The
breast wheel has water entering the wheel more or leas radially, filling the buckets and then again being
dumped near the bottom of the wheel. Typical efficiency values vary from as low as 15% for the
undershot to well over 50% for the overshot with the breast wheel in-between.
We shall concentrate on the overshot wheel as being the most likely choice to give maximum power
output per dollar cost, or per pound of machine, or per manhour of construction time based upon expected
efficiences. Mitigating against this choice is the need for a more complex earthworks and race way with
the overshot wheel where the water must be guided in at a level at least as far above the outlet as the
diameter of the wheel.
The undershot wheel, of course, may be merely set down on top of the stream with virtually no
preparation of raceway necessary. But in many streams the rise and fall with heavy local rainfall is
spectacular, so flood protection would be a major consideration for any type of device.
The simplest flood protection is a channel leading from the river to the installation, with inlet to the
channel controlled to keep flood water in the main stream. Since a diversion channel would probably be
required anyway, the odds are very good that a suitable location to employ an overshot wheel can be
found for most installations. In the event that the overshot installation is impossible, the undershot wheel
straddling the diversion channel is simple to use.
Another consideration which makes the overshot wheel attractive is the ease with which it can handle
trash in the stream. First, the water shoots over the wheel and so trash tends to get flung off into the tailrace without catching in a bucket. Secondly, there are not usually the tight spaces between race and wheel
in which trash can jam. Somewhat closer fitting arrangements are required with breast and undershot
wheels to get good efficiency.

III. LIMITATIONS - ADVANTAGES AND DISADVANTAGES


The wheel is a slow speed device limited to service roughly between 5 and 30 rpm. Consequently this
limits its usefulness as a power source for electricity generation or any other high speed operation
because of the step up in speed required. Although not a great problem from an engineering viewpoint,

adequate gearing or other speed multiplying devices involve increasing complexities in terms of money,
potential bearing problems, and maintenance.
The slow speed is advantageous when the wheel is utilized for driving certain types of machinery already
in use and currently powered by hand. Coffee hullers and rice hullers are two which require only
fractional horse-power, low speed input. Water pumping can be accomplished at virtually any speed.
Slow speed output of a wheel cannot of course, directly power a centrifugal or axial pump.
The positive displacement "bucket pump" or suction lift pump already in use in various villages normally
operates at well under 100 cycles per minute and can be adapted for use in conjunction with a wheel at
slow speed. This of course, has been done for hundreds - maybe thousands - of years elsewhere.
Devices of this type have relatively low power output capability. The power output depends upon the
dimensions of the wheel, the speed and the useable flow rate of water to the wheel. As an example, a
reconstructed breast wheel installed in a museum in America of 16 ft. outside diameter and with bucket
depth of 12 in. operating at 7 rpm, with flow rate of 28 cubic feet of water per second had an estimated
power output of 18.5 hp (14 kw) (calculated at an efficiency of 100%).
Actual output on that wheel has not been measured but would be less than 10 hp (7.5 kw). A 3 ft. OD, 1
1/2 ft. vide model constructed by the author is in the fractional horse-power range.
Already mentioned once, it is worth emphasizing that a useable water wheel can be built almost anywhere
that a stream will allow, with the crudest of tools and elementary carpentry skills.

IV. THEORETICAL CONSIDERATIONS


A. Stall Torque
The stall torque capacity of the machine, ignoring the velocity effect of the water impinging on the stalled
buckets, is easily calculated by a simple summation of moments about the shaft due to the weight of
water in each filled or partially filled bucket. Obviously this will depend in part on the amount of spillage
from the bucket which in turn depends on bucket design. Bucket configurations used in the 18th and 19th
century varied depending on the skill of the builder.
They were empirically determined on the criterion of maximizing torque by maximizing water retention
in the buckets while recognizing that optimum design on this criterion also required increased
construction complexities. Buckets of shape shown schematically in a side view, Fig. 1, were used for
overshot and breast configurations. The straight sided buckets are less efficient but simpler to construct.
The width of the bottom of the bucket was typically 1/4 of the width of the annulus where that
configuration was chosen. Purely radial buckets were used in undershot wheels.

It is convenient to use three of the wheel's dimensions for calculation of the torque capacity of the wheel:
the outside radius, r; the wheel width, w, i.e., from side to side; and the annulus width, t, defined as
t = (outside diameter - inside diameter)/2.
See Fig. 1.
The ratio of the annulus width, t, to the outside radius, r, is important to wheel design as there are
practical limits to the useful values which may be employed. In this paper only ratios 0.05 t/r < 0.25 are
considered. For smaller ratios, the potential output per foot of diameter of the wheel is considered too low
to be practical. For larger values, the buckets become so deep that there is insufficient time to fill each
one as it passes under the race exit.
Also, since the torque and power depend upon having the weight of water at the greatest possible distance
from the wheel axis, increasing annulus depths increases total wheel weight faster than it increases power
output. The result is that if more power is needed it is better to increase the O.D. than to increase the
annulus width to values exceeding t/r = 0.25.
In this way the wheel weight and the structural components to support that weight remain economically
most advantageous for a given power output. Historically, wheels have tended to have t/r values around
0.1 to 0.15.

Upper limits on wheel width have tended toward approximately 1/2 the O.D. because of structural
problems with wider wheels.
It can be estimated that the overshot wheels operate with the equivalent of approximately 1/4 of the
buckets full. That is, the total weight of water doing useful work on the wheel is 1/4 of the total that
would be contained in an annular solid of dimensions the same as the O.D., I.D. and width of the wheel.
The actual weight distribution of the water is as shown schematically in Fig. 2a because of spillage from
the buckets as they approach the tail race. If we assume the water is concentrated in the annular quadrant
shown in Fig. 2b, the stall torque can be estimated more easily. A suitable correction factor could be
applied to account for actual bucket design, if that refinement were considered necessary.

Figure 2. Schematic view of water distribution on wheel. A) Approximate actual static distribution.
Crosshatching indicates water remaining in buckets. B) Approximation of weight distribution used for
calculations in this paper.
Results for wheels of various dimensions are given in Table 1.

TABLE I
Stall Torque Per Foot of Width (ft.lb) (No Allowance for
Thickness; .05 < t/r < 0.25 only)
Outside Diameter (ft.)
Annulus Width (in)
3
4
6
8
20 40
95
2
30 55
130
3
40 70
160
4
95
240
6
300
8
10
12
16
20
24
Blank spaces are left for t/r ratios outside the limits given above.

Volume Consumed by Bucket Wall

10
235
300
435
555
655
773

14
363
485
695
905
1080
1270

20

1435
1850
2220
2620
3300
4000

2910
3840
4660
5560
7190
8750
10100

Experience has shown that many non-technically trained users of this information will be more confident
of their ability to use data given in tabular than in graphical form. Both will be presented here when
appropriate.
B. Power Output
Power output is the product of the torque on the output shaft and the rotational speed of the shaft. On the
assumption that there is sufficient inlet water flow to keep the buckets full, thereby keeping the torque
constant, the power output increases linearly with speed. In a location where there is virtually an
unlimited inlet water supply, this calculation will give an upper limit to the power output that can be
expected.
The horsepower output per rpm per foot of width is shown in Table II.

TABLE II
Horsepower Output for a Constant Torque Wheel per. RPM per Foot of Width
Outside Diameter (ft.)
Annulus Width (in)
3
4
6
8
10
14
0.0042
0.0072 .018
X
X
2
0.0053
.011 .024 .044
.0690
3
0.0070
.013 .030 .057
.0920
4
.018 .045 .082
.152
6
.057 .105
.171
8
.123
.204
10
.137
.238
12
16
20
24

20
X
X
X
.269
.347
.416
.490
.618
.750

X
X
X
0.545
0.720
0.872
1.05
1.35
1.64
1.90

At any given speed, the actual horsepower output will be the Table II entry appropriate to the size wheel
used times the actual speed in rpm times the width of the wheel in feet.
The water power input is the maximum power which the wheel could achieve if it were 100% efficient. It
is calculated as the product of the water's specific weight, the volume flow rate, and head and is given in
Table III for comparison. This entry is also in horsepower per rpm per foot of width of wheel. The
volume flow rate is that required to keep the buckets full and is given in Table IV.

TABLE III
Water Input Horspower to Wheel per RPM per Foot of Width to Maintain Constant Torque.
Outside Diameter (ft.)
Annulus Width (in)
3
4
6
8
10
14
20
.0086
.0154
.0364
2
.0118
.0226
.0510 .092 .143
3
.0158
.0268
.0646 .119 .189
4
.0416
.0980 .176 .280 .568
1. 11
6
.128 .230 .368 .740
1.52
4
.277 .448 .900
1.85
8
.316 .537 1.08
2.23
12
1.39
2.94
16
1.73
3.64
20
4.26
24

TABLE IV
Flow Rate In Imperial Gallons set RPM per Foot of Width Of Wheel Required to Maintain
Constant Torque
Outside diameter (ft.)
Annulus Width (in.)
3
4
6
8
10
14
20
10
13
20
2
14
18
28
59
48
3
18
24
37
50
62
4
34
54
74
93
123
190
6
70
96
123
175
254
8
115
149
216
312
10
137
177
255
373
12
342
488
16
404
600
20
706
24
As in Table 1, no allowance has been made for the volume occupied by the bucket wall thickness. This
can be corrected for later if desired. The head is assumed here to be the diameter of the wheel. The lower
edge of the wheel is the highest elevation permission for tailrace water without interfering with the wheel
and is a logical datum.
Inlet raceways are seldom found with a significant slope so that velocity effects of raceway water are
small. It seems sufficiently accurate to estimate the inlet elevation as the top of the wheel. Any error
thereby introduced will be on the conservative side anyway.
Theoretical efficiency values for the wheel using the assumptions adopted so far can be found by taking
the ratio of the power output from Table II and the corresponding power input of Table III. These values,
for the water weight distribution assumed before, are about 50% for the narrow annulus wheels and drop
to just under 45% for the wider annulus wheels. As mentioned previously, a well designed and
constructed wheel will give efficiencies better than this.
This comparatively modest value is primarily the result of not considering the effect of the water still in
the buckets below the horizontal centerline. It reflects the fact that the simplifying assumption that the
buckets remain full half way down the wheel and suddenly dump all their water is not accurate. That
inaccuracy is tolerable because 1) it makes the analysis so simple and 2) it gives slightly conservative
figures for power so that almost every reader will be assured of getting sufficient power even from wheels
of relatively amateurish construction.
When the water flow is less than the required to fill each bucket completely as may be the case for a
stream of limited size, the power characteristics are altered in that the torque now is a function of speed.
Using the assumption of one annular quadrant working, but not full, the volume of water, V, in the
quadrant is
V = Q/4N
where
Q = volume flow rate ([ft3]/min)
N = speed (rpm)
The weight of water in the annular quadrant at any speed is then pgV where
p = density of water
g = gravitational acceleration
With units in feet, pounds, and minutes, the horsepower to be expected from this annulus working is
hp = 2[pi] NT
-------33,000
where T =
pgV[bar]x

pgQ[bar]x
--------4N

[bar]x is the distance to the centroid of the annular quadrant from the rotation axis. It is equal to average
diameter. [D.sub.av], of the annulus divided by [pi].
Therefore
hp =

2[pi]NpgQ[Dav]
-------------4[pi]Nx33,000

pgQ[Dav]
--------66,000

The power is independent of the speed. The efficiency is the same as calculated previously. It is because
the output power is a function of the average diameter, that the efficiency drops off for wide annulus
wheels of a fixed outside diameter. Potential power output from a wheel operating under the conditions of
constant flow may be estimated most easily by the equation for water input power, assuming 50%
maximum efficiency and head equal to the outside diameter.
Power under constant flow conditions for various diameter wheels is shown in Table V for likely
attainable flow rates. The values shown are for t/r = 0 and should be corrected by multiplying Table
entries by factors as shown at the bottom of the table for various practical t/r values. The author's
prototype with t/r = .17 tested at approximately 150 gpm, gave output power of approximately .06 hp in
reasonable agreement with the values predicted in Table V.

TABLE V
Estimated Maximum Output Horsepower from Wheel for Constant Input Water Flow Rate
Condition (based on 50% efficiency of wheel)
Flow rate (gallons per minute)
Outside diameter (ft.)
100
200
500
1000 2500 5000 10000
30000
0.045
0.091
.23
3
0.060
.12
.30
.60
4
.091
.18
.45
.91
2.27
6
.24
.61
1.21
3.03
6.1
5
.30
.76
1.52
3.79
7.6
10
.42 1.06
2.12
5.30
10.6
21.2
14
1.51
5.03
7.58
15.2
30.3
91.0
20
Correction factors for various values of t/r: to be multiplied by above values to give correct power ratings
0.05 0.1 0.15 0.20 0.25
t/r
0.98 0.96 0.94 0.92 0.90
correction factor
Blank spaces are left where flow rates are impractical for the wheel size given. Upper bounds to practical
flow rates for various wheel sizes are found by multiplying the entry from Table 1 by the practical upper
limit of speed and width for the O.D. and are shown in Table VI. Lower bounds are subject to
considerably more guesswork.
On the assumption that it would be uneconomical to construct a wheel of width less than 1 ft. and to
operate it at less than 25% capacity (completely arbitrary choice) for the speeds quoted in Table VI the
useful lower bounds may be estimated. These are indicated by blank under the 100 gpm and 200 gpm
columns in Table V.
TABLE VI
Upper Limits on Useable Flow Rates for Various Size Wheels in Imperial Gallons Per Minute
(assuming wheel width = 1/2 (O.D.) and peripheral velocity is 5 ft/sec.)
Outside Diameter (ft.)
3
4
6
8
10
14
20
RPM at 5 ft/sec peripheral velocity
32
24
16
12
10
7
5
Annulus Width +(in.)
500 625
1000
2
700 900
1400
1900
2500
3
900
1150
1800
2400
8000
4
1650
2600
3500
4500
6000
9500
6
3400
4500
6000
8500
12000
8
5500
7500
10500
15500
10

6500
9000
12500
18500
12
17000
24000
16
20000
30000
20
35000
24
The upper limit to the speed at which the wheel will operate depends primarily upon the rate at which the
wheel slings the incoming water off so that it is not utilized. This depends primarily upon the speed and
radius of the wheel and secondarily upon the bucket configuration and its relation to the inlet water.
The figures quoted in Table VI are based on the rule of thumb peripheral velocity of 5 ft/sec. With
smaller wheels this is a bit high, based on prototype tests. With the larger wheels the peripheral velocity
may be as high as 8 ft/sec.
In summary, the type of power vs. speed curve that one can expect from a water wheel is as follows for
fixed flow rates: Linearly increasing from zero speed up to the speed at which the buckets can no longer
be completely filled by the prevailing flow, then constant up to the speed at which significant amounts of
water are rejected from the wheel by slinging action, thereafter decreasing in proportion (roughly) to the
square of the speed.
C. Bucket Design
The optimum bucket design is taken to be that which produces the greatest torque on the wheel shaft. The
upper limit to this condition is that the buckets fill completely at the top, carry the full water weight with
no spillage to the bottom and dump their loads there. There is not a practical method of achieving this
maximum. With fixed buckets, the best we can do is minimize spillage from the buckets as they travel
from the top, where they are filled, to the bottom where they should be empty (so as to limit losses
incurred by carrying water up the back side of the wheel).

There are broadly two styles of bucket as shown in Fig. 1. In the straight sided bucket the limits on the
angle the bucket makes with the tangent at the O.D. or I.D. (See Fig. 1) are from tangential (0) to radial
(90). With tangential buckets, the filling process is slow at the top because of the very shallow angle
with respect to the incoming(nearly horizontal) water.
Furthermore the emptying process at the bottom is not complete until after the bucket passes bottom dead
centre. This carries some water up the back side and consequently reduces the efficiency. At the other
extreme, radial buckets are nearly empty by the time they have gone 1/4 turn from the top because the
bucket wall is then horizontal.
We can estimate the optimum angle by assuming that the greatest effect will be due to the bucket whose
weight is acting at the greatest distance from the shaft. By drawing buckets of various angles we can
estimate, graphically, the optimum. While the tangential bucket carries the greatest amount of water, its
centroid distance is not a maximum
The maximum occurs at a bucket angle (to the tangent at the I.D.) of about 20. While the amount of
water still retained at 90 after top dead centre by this bucket shape is about 20% less than for the
tangential bucket, the loss is compensated for in the early filling and early emptying.
Especially on emptying, the 20[degrees] inclination is a major factor since the length of the bucket
(distance from I.D. edge to O.D. edge) is more than 30% shorter than the tangential bucket. With a 30bucket, the weight carrying capacity at 90 after top dead centre is down to about 65% of the tangential, a
figure which is so low that it cannot be compensated for by the secondary effects on efficiency such as
filling and emptying.
This graphical technique, while of no additional value in designing any individual wheel, also shows that
the assumption of the distribution of water over an upper quadrant is a reasonable one for estimating
torque.
It is recommended that the bucket wall angle be kept between 200 and 250 to the I.D. tangent.
The use of flat bottomed buckets does not significantly change the water carrying capacity for wall angles
of 20. The purpose is to decrease the distance the water must travel to empty the bucket. Its use is
increasingly beneficial at large t/r ratios but the builder must accept that the construction is somewhat
more complicated than that of the straight sided bucket.
Bottom widths should be approximately 1/4 of the annulus width, t. This will cut 25% off the side width
with the attendant saving in travelling distance to empty the bucket. The significance of this is that less
water is carried up the back side of the wheel. Any water carried up the back side lowers the efficiency. I
cannot give figures for the improvement of efficiency using flat bottomed buckets but it seems hard to
imagine as much as ten percentage points.
Historically, bucket shapes have varied considerably. They were, as far as can be determined, chosen
emperically. (In a historical sense this is a euphemism for "arbitrarily" or "by educated guesswork"). By
the time engineers, rather than carpenter-craftsmen, were considering the problem the water wheel's
usefulness was already on the decline).
Even in relatively recent manuals for construction, circa 1850, while wheels were still in general use in
the U.S., bucket side angles of 45 were recommended - a choice which can easily be shown to be less
efficient than smaller angles. The 20 - 25 figure is, however, in close agreement with the design of two
wheels that I know are still in use in the U.S.
The number of buckets to use depends upon the volume consumed by the bucket wall material. The ideal
wheel has closely spaced buckets of very thin wall thickness. A reasonable figure to design by is that not
over 10% of annular volume should be consumed in bucket material. Typical values for the size wheels
discussed here would be 25 - 30 - 1/4 in. thick buckets on a 3 foot wheel and 50 - 1-1/4 in. thick buckets
on a 14 foot wheel.
D. Bearing Design
The wheel itself has only one rubbing or sliding part subject to wear, viz. the bearings upon which the
axle is supported. Standard bearing design is covered in almost any machine design text. In the
manufacture of such a device as is discussed here, the value of such standard" bearings is questionable.
Fully weather-proofed ball or roller bearings are too expensive and complicated to satisfy the initial
criteria.
Bronze bushings with suitable shaft material would be satisfactory but lubrication and replacement both
present problems. The use of wooden bearings is, I think, the best alternative for several reasons:

1. Simplicity of manufacture with local skills.


2. Availability of replacement parts.
3. Negligible cost.
Wooden bearings are used commercially for such applications as washing machine wringer bearings
under conditions simulating those proposed for the wheel. Rock maple, lignum vitae, and various species
of oak are used commercially, but when these are not native to the country of intended use, substitutes
may be found.
Among woods with fairly widespread distribution, others which may reasonably be expected to be
satisfactory are beech and red mangrove. Forestry departments, when they exist in a country are generally
in a position to make useful suggestions.
In the absence of any specific knowledge, the general rule is "the harder, the better".
An estimate of allowable loading based on experience with commercially available wooden bearings
would be around 75 psi (for oak) to 150 psi (for lignum vitae) for orientations with the sliding surface
parallel to the grain and about 150 to 300 psi respectively for end grain use.
If the wood used has strength and density properties comparable to those mentioned above, it is likely
that safe loading would be about 100 psi parallel to the grain and 200-250 in end grain usage. It remains
to be seen what the wear resistance at these pressures will be, but structurally the figures given can be
used with confidence.
Length to diameter ratios of bearings in this application would reasonably be expected to be about unity
and on that basis the sizes of the bearings can be estimated for wheels operating at maximum output. An
allowance for the weight of the wheel itself is made on the basis that the volume of wood required is
approximately equal to the volume of water carried at stall and that the specific gravity of wood operating
constantly in water is about unity.
Table VII shows the approximate weight on each bearing per foot of width of wheel. Total weight carried
on each bearing is then the product of the Table entry and the width of the wheel in feet. This of course
assumes that the wheel is simply supported at each end of the shaft and does not allow for additional
loads imposed by the attached machinery.
It is important that significant loads due to the Table VII values for the purposes of determining bearing
size from Table VIII for the side of the wheel where the machinery is attached. In this event the bearings
will apparently need to be of different sizes. In practice, unless the indicated sizes are very different, we
usually make both the size indicated by the larger load. Thus one is really longer than it needs to be.
Bearing diameters required to support the various loads are given in Table VIII calculated on the basis of
100 psi in parallel useage and 200 psi for end grain useage and L/D = 1. Values are given to 20,000 lb. to
allow for the largest reasonable bearing loads.

TABLE VII
Approximate Weight Carried by Each Bearing Excluding Loads Due To Attached Machinery (per
foot of width of the wheel) (lb.)
Outside Diameter (ft.)
Annulus Width +(in.)
3
4
6
8
10
14
20
24 32 SO
2
35 47 70
95
120
3
44 60 89
125
160
4
86 140
185
235
335
470
6
180
240
305
440
675
8
290
370
530
765
10
330
445
635
920
12
820
1215
16
1020
1500
20
1760
24

TABLE VIII
Minimum Bearing Diameter Required for Various Loadings (in.)
Load (lb.)
Parallel Useage

100 200

500

1000

2000

5000

10000

20000

2-1/4

3-1/4

4-1/2

10

14

1-1/2

1/2 1
1-3/4 2-1/4 3-1/4 5
7
10
End Grain Useage
These bearings are assumed to be steel on wood. In the likely event that, especially in larger sizes, the
bearing is considerably larger than the required shaft size, a "built up and banded" bearing may be used.
A wooden cylinder is built onto the shaft at the bearing location such that the cylinder O.D. is the
necessary size.
Then steel bands are bent and fastened to the cylinder. The criterion for design in this case is that the
product of the diameter and the total width (sum of the individual widths) of the bands equals or exceeds
the square of the entry in Table VIII for the corresponding load and grain orientations.
If it is possible to arrange for and be certain of, suitable maintenance, a steel shaft in bronze bushings
mounted in commercial plummer blocks (available from hardward suppliers) is probably the best choice.
Proper alignment may be a minor problem but is usually fairly easy to overcome. This choice involves
additional initial expense and is justified only if maintenance can be guaranteed regularly and frequently.
E. Shafts
Shafting may be wooden or steel. The diameter is of course dependent upon which material is used and
the dimensions of the wheel. Minimum permissible shaft diameters d, may be estimated from the
equation for stress for solid metal shafting
[d3] =
16 [square root][M2] + [T2]
----------------------------[pi]S

In this equation M is the maximum bending moment occuring where the wheel sidewall attaches to the
shaft. It can be estimated as the product of the bearing load (entry in Table VII for the appropriate wheel)
and the distance from the wheel side wall to the centre of the bearing. In the interest of keeping the shaft
as small as possible, it is therefore desirable to locate the bearings as close to the side of the wheel as
possible.
(Note that in most cases, it is not critical to include the additional machine load on the bearing, discussed
in connection with the use of Table VIII. It must be included only when the external machine load times
the distance along the shaft from the point of application of the load is larger than the bearing load from
Table VII times the distance along the shaft from the bearing to the point where the wheel is attached.)
T is the torque acting on the shaft and a conservative estimate is found from Table I. S is the allowable
shear stress of the metal. It is around 15,000 psi for ordinary steel shafting and pipes. (13,000 is used in
the example in Appendix 1.)
For solid wooden shafts two equations are used and the larger diameter of the two results is chosen as the
diameter of the shaft.
[d3] = 16T
---[pi]S
[d3] = 32M
---[pi]B

where S, T and M have the same meaning as before. However, the value of S is typically 150 to 300 psi
for hardwoods. B is the allowable bending stress and has a value of about 1500 psi for typical hardwoods.
If wood is used it must be sound and free from longitudinal cracks.
For hollow shafting like a pipe, the equation to determine the outside diameter is:

[d3] =

16[square root][M2] + [Ts2]


--------------------------[pi]S(1 - [k4])

where K = Ratio of inside to outside diameter.


The values of O.D. and I.D. are standardized for pipes. For bearing loads tabulated in Table VIII, on the
assumption that the centre of the bearing is 1 foot from the edge of the wheel, the standard pipe sizes
shown in Table IX should be satisfactory. Table IX automatically allows for torque that would be
reasonable to expect from a wheel of such a size that the bearing load would be given in Table VIII.
The values are approximate only since exact values cannot be given until all the details concerning the
loads due to the attached pump or machine are known. The values given should serve as a guide only and
the final decision should be checked against the equation to be sure. When making substitutions, in
assembly, of one pipe size for another, it is permissable to use larger pipe than shown in Table IX but not
smaller pipe.
TABLE IX
Minimum Standard Pipe Sizes for Use as Axles with Bearings at 12 inches from Wheel Edge
Bearing load (lb)

100

200

500

1000

2000

5000

10000

1"
1 1/2"
2 1/2"
3"
4"
6"
8"
Pipe Diameter (in)
Comparing these figures with the required bearing diameters of Table VIII, it is obvious that when using
pipe or solid steel shaft, the bearing will need to be of the build up type when using wooden bearings. An
alternative is to use a shaft whose size is selected according to the needs of the bearing size. It will be
much stronger (and heavier) than necessary but may save some work.
With wooden shafts, the required shaft diameter will usually exceed the required bearing diameter and
then one has the choice of reducing the shaft diameter at the bearing location (but only there) or of using
larger bearings. In either case the shaft must be banded with steel, sleeved with a piece of pipe or given
some similar protection against wear in the bearing.
F. Minor Considerations
We have considered all the major theoretical aspects of selection of sizes etc. to meet specific
requirements. All have been based on an assumed efficiency of 50% - a figure which is readily achievable
in practice with an overshot wheel. There is one minor consideration over which the design/builder has
control which may affect the effiency slightly. The raceway exit should put water onto the wheel slightly
before top dead centre. The exact location is a function of
1. flow rate and raceway inclination which affect the inlet water velocity; and
2. the bucket sidewall angle and peripheral velocity which affects how smoothly the inlet water comes
onto the wheel.
Exact calculations hardly seem justifiable for a machine which by its very nature is as crude and
(relatively) inefficient as this. Let it be sufficient that the designer-builder get the water in approximately
tangent to, and at the top edge of, the wheel.

V. PRACTICAL CONSIDERATIONS
A. Materials
Most wheels are wood, of course, though they need not be. Among the considerations for selection of the
proper material are the ease of working, cost, availability and durability. The average carpenter can make
a proper choice on all these except perhaps the latter. Forestry departments in many countries can provide
this information on potentially useful species. Others which would probably be suitable are mentioned in
the section on bearing design.
Builders of water wheels may naturally consider a "marine" plywood as a likely material. It is convenient
to work with but the quality varies widely around the world. Because even the best grades have a doubtful
durability when operating continuously in water unless painted, plywood should be chosen only when it
can be well cared for or when a relatively short life is anticipated
Regarding the framework to mount the wheel on, bamboo might seem a logical choice in many countries
but the durability is such that it probably would require more long term care and replacement than other

materials. The species listed for the bearings in section IV D are all fairly durable under constantly wet
conditions and should be the first to be considered.
B. Construction Techniques
Any person sufficiently skilled to build a water wheel will probably also be sufficiently knowledgeable to
work out most of the construction details. This manual is intended to give the engineering groundwork
necessary to select the proper overall size of wheel to meet a given need and to make sure that prevailing
water supplies are, in fact, adequate. However, a few general suggestions may help the reader avoid some
pitfalls.
Attachment of the wheel sides to the shaft, whether the sides are spoken or solid, can be accomplished in
many ways. If a steel shaft is used, a thin flange plate can be welded to the shaft (if such facilities are
available) and this greatly facilitates the attachment. With a solid side plate there is no further problem
but if the spokes are used, the bending in the spokes at the flange must not be so great as to break the
spokes.
The spokes should be attached to the flange with two or more bolts and the distance required between the
bolt holes to support bending varies with wheel diameter and the rigidity of the spoke/wheel joint. For a
flexible joint the required a distance would be approximately 1/10 to 1/12 of the outside diameter of the
wheel.
For example, on a 12 foot wheel, when using radial spokes attached to a flange by 2 bolts and to the
wheel side plate (annular ring) by one, the flange bolts should be about a foot apart on each spoke.
Alternatively if the spokes are quite rigid and firmly attached to annular ring of the wheel as with 2 or
more bolts, the bolt hole separation can be reduced to 1/20 of the diameter of the wheel at the flange.
A simple spoke arrangement to use is pairs of spokes, (one spoke of each pair on each side of the shaft)
crossing at right angles to make a shape like the tic-tac-toe or noughts and crosses symbol. The wheel
axis runs through the centre square and the extremities of the lines are attached to the wheel annulus.
Any glue used should be highest quality waterproof glue for obvious reasons. Resorcinol glue is probably
the best choice.
Bucket attachment to the side wall may be made by either grooving the side wall to receive the bucket
edge or by attaching strips to the inside of the side wall to fasten the buckets to. There is an advantage to
the annulus shape of side wall in that the inside of the bucket is accessible from the I.D. This makes
closing off the inside of the bucket simpler because the necessary pieces can be inserted through the I.D.
With solid sidewalls, the buckets must be made complete and non-leaking before the sidewall is attached.
This is by no means impossible but may be more difficult.
If a solid sidewall is used, holes should be drilled adjacent to the bucket bottom into the space between
the bucket and the haft to let any leakage water out. A solid side wall would not commonly be used.
Spokes offer several advantages.
Numerous books are available to give helpful hints on various construction techniques for the truly
amateur builder.
C. Maintenance
The wood used may be painted or varnished for a protective coating. This will obviously extend the life
of the wheel. Periodic repainting, if desired, can be carried out. The decision on painting must be made on
purely economic grounds. If a very durable wood has been used initially, painting is a luxury. If a
somewhat less durable species is used, painting is probably cheaper and easier than early replacement or
repair of the wheel.
The only major maintenance problem is in bearings. Generous allowances have been made in the figures
in Table VIII but the bearing will still ear. This will drop the wheel from its original position. Shimming
under the bearing block will compensate for this. Bearing replacement, when the block is completely
worn through is a simple matter.
Lubrication is totally unnecessary with lignum vitae or commercially processed maple, if available. With
the other species, we can not make such a flat statement. Generally speaking the bearing should be made
from the hardest wood available and lubricated as needed. Oils and grease in small amounts will probably
do no harm and may slow the wear rate. Pig grease and tallow would certainly be harmless and might
help.

PART TWO: APPLICATIONS


I. WATER PUMPING
A. Pump Selection
The only type of pump which is reasonable to use at the slow speed of the wheel is a positive
displacement pump. They are called by various names such as bucket pump, lift pump, piston pump,
windmill pump and occasionally even simply by brand name such as "Rocket" pump. Numerous models
are available commercially and vary in cost from a few dollars for small capacity pumps to several
hundred for high capacity, high head, durable, well manufactured units.
Units can be manufactured at low cost in the simplest of workshops. Details are given in Appendix II.
Such pumps may vary in bore size, stroke length and head capacity. There is a practical limit to the speed
at which they can operate. This is usually above the frequency of the fastest of wheels. A frequency of
speed multiplier such as a multi-lobed cam or a gear set may be used, but these more complicated pumps
and mechanisms, while increasing the efficiency of the pumping process, contravene the criteria of
Section II, Part One for simplicity and will not be discussed. We will discuss only very simple pumps.
Even with simple single or double acting pumps there are certain problems. one single acting pump
attached to the wheel will cause speed surges on the wheel because of the fact that actual pumping takes
place only half the time. The other half is spent filling the cylinder. During this filling stage considerably
less wheel torque is required than when actually pumping. The speed surge can be partially overcome by
using
1. two single acting pumps 180 out of phase so that one of the pumps is always doing useful work;
2. a double acting pump which has the same effect as 1. but is built in one unit; or
3. best of all two double acting pumps 90 out of phase.
Such use of multiple simple pumps will also improve the overall efficiency of the system. (In general one
unit can be attached easily to a crank at each end of the wheel shaft).
There are pressure variations in the delivery line which depend on several factors. As long as the peak
pressures do not exceed the capacity of the pump and related mechanism, nor stall the wheel, such
variations will cause no harm. The pressure peaks can be damped with an air chamber in the delivery line
or smoothed by using two or more simple pumps as mentioned in the preceeding paragraph.
The possibilities are so numerous and the details sufficiently complex that they cannot all be included
here. A pump expert or pump design manual should be consulted if the design ideas given here seem
insufficient for the user's needs.
In general the pressure peak will be a function of the peak piston velocity, the pump bore size, the
delivery pipe size, the length of the delivery pipe and the type of pipe used. When speaking of pump
performance and design requirements, the term "head" is encountered often. It is a means for visualizing
the fluid pressures involved in the pump or attached pipes.
It means the height of water in a vertical pipe necessary to produce, at the bottom of the pipe, the pressure
being referred to. The pressure is an actual system will not, in general, be produced just by a static
column of water but it will be the same as if it were. It's just a handy shortcut often used by fluids
engineers. The head The head required at the pump outlet will be made up of two main components:
1. the actual change in elevation to the delivery pipe exit, i.e. the (vertical) height of the hill; and
2. friction loss in the pipe which is given by the equation:
friction loss = f

L V
--D 2g

where f = friction factor obtainable from handbooks or Table X.


L = length of pipe
D = inside diameter of pipe
V = velocity of water in the pipe
g = gravitational acceleration
(Note: Units for dimensions must be consistent. See Appendix I for an example of the use of this
equation).

TABLE X
Estimated Friction Factors for Cool Water
Water Velocity (ft/sec.)
1
5
10
.045
.040
.038
Old Iron Pipe
.030
.023
.021
New Iron Pipe
.025
.017
.015
Plastic Pipe
It is evident that this becomes a major factor in very long pipes, in small diameter pipes, or with high
velocities. The water velocity in the delivery pipe is a function of the peak pump piston velocity and the
ratio of the pump bore size and the delivery pipe size. Peak piston velocity for pumps attached directly to
the wheel is given in Table XI for various strokes and wheel speeds.
From Table XI, the delivery line velocities can be estimated simply by multiplying the Table XI entry by
the ratio of the pump bore area and the delivery pipe area. That is, piston velocity times piston area =
water velocity in delivery pipe times pipe bore area.
As a rule of thumb, this resulting delivery pipe velocity should be a maximum of 10 ft/sec. in short runs,
and even smaller for very long pipes. The peak head required of the pump will be the sum of the two
different heads mentioned, i.e., elevation change plus friction loss head.
The bore size (piston area) and peak head occurring during pumping will determine the force required at
the pump rod since force on an area is the product of the area and the pressure acting on that area. Figures
for force at the rod are given in Table XII. No allowance is made for rod diameter so the figures given are
conservative. Bore sizes quoted are commercially available.

TABLE XI
Peak Pump Piston Velocity (ft/see) for a Pump Rod Attached Directly to a Crank on the Wheel
Stroke (in.)
Wheel Speed (RPM)

2 1/4

10

12

0.048

0.087

0.129

0.172

0.216

0.2 60

.059

.104

.156

.208

.259

.310

.078

.138

.207

.276

.345

.414

10

.097

.173

.259

.345

.432

.518

12

.117

.208

.312

.416

.520

.624

IS

.147

.2 60

.390

.520

.650

.780

20

.195

.345

.518

.690

.865

1.04

TABLE XII
Peak Force on the Pump Rod of a Piston Pump Required for Various Bores and Heads (lb.)
Peak Force on the Pump Rod of a Piston Pump Required for Various Bores and Heads (lb.)
Peak Head (ft.) change in elevation and friction loss
Pump Bore (in.)
50
100
200
300
400
500
30
60
110
370
220
280
1 1/4
40
80
160
240
320
400
1 1/2
60
110
220
320
430
540
1 3/4
70
140
270
420
560
700
2
110
220
440
660
880
1100
2 1/2
185
370
740
1120
1480
1850
3 1/4
315
630
12 60
1890
2520
3150
4 1/4
These figures are required to design such parts as clevis pins (if used) and to determine that, if the pump
rod is attached directly to the wheel, that the crank arm length times the entry in Table XII does not
exceed the torque capacity of the wheel as given by Table I.

Of course, if levers or other torque/force multiplying devices are used, appropriate calculations at the
wheel can be made. The force at the pump rod still remains as given in Table XII. The velocity given in
Table XI must be adjusted for the change in crank arrangement.
Additionally, if the line is very large so that a large mass of water must be accelerated on each stroke, the
inertial forces can become greater than the pressure forces. The inertial forces can be estimated with the
aid of Tables XIII and XIV.

TABLE XIII
Volume of fluid in various sized delivery pipes (ft.)
Pipe length (ft.)
Nominal pipe size
50 100 200 500 1000
.3
. 6 1.2
3
6
1"
1.16 2.32 4.65 11.6 23.2
2"
2.46 4.91 9.82 24.6 49.1
3"
4.38 8.78 17.50 43.8 87.5
4"

TABLE XIV
Inertial force (lb.) per inch of stroke for various volumes of fluid at various pump cycles speeds
Volume of Fluid in delivery pipe (ft3)
Pump Cycles per Minute
.5
1
2
5
10
SO
100
.133 .266 .533 1.33 2.66 13.3 26.6
5
.577 1.14 2.29 5.77 11.4 57.7 114
10
1.20 2.40 4.80 12.0 24.0 120
240
I5
2.14 4.27 8.33 21.4 42.7 214
427
20
3.31 6.61 13.2 33.1 66.1 331
661
25
4.78 9.65 19.1 47.8 96.5 478
965
30
This inertial force is at its peak just as the piston starts its pumping stroke. At this time the friction loss is
zero because the delivery pipe velocity is zero. Hence the total rod force at the start of the stroke will be
equal to the force due to the static head plus the inertial force. It should be compared with the rod force
when the friction loss is a maximum and the components designed to withstand the larger of the two.
We can calculate the power required to accomplish pumping under various conditions of head, flow rate
and pump type. These figures are given in Table XV for steady flow and are adjusted for unsteady flow
explained below.
This is the theoretical minimum power input required to the pump under steady conditions.
Under the unsteady conditions of a piston pump, to estimate the water wheel power capacity required,
multiply the table entry by 2 1/2 for one single acting pump, by 2 for one double acting pump or two
single acting pumps 180[degrees] apart or by 1.5 for 2 double acting pumps 90 apart. This will give an
estimate of the size of wheel and flow rate required to the wheel.
As mentioned near the beginning of this section, there will be speed fluctuations in the wheel which may
be pronounced in smaller wheels working near their capacity. This is no particular disadvantage so long
as the stall torque capacity of the wheel exceeds the minimum torque necessary to keep the pump
moving. The magnitude of the fluctuations decreases with double acting or multiple pumps installations
and where the mass of the wheel is such that a flywheel action begins to take place.
The volume pumped per stroke varies slightly with the design of the pump and with the bore and stroke
sizes. One commercial manufacturer quotes figures which can be taken as representative.
These are given in Table XVI.

TABLE XV
Horsepower Required for Water Pumping at Various Flow Rates and Heads (both assumed steady)
Total Head (ft.)
Flow Rate (imp. gal/hr.)

50

100

200

300

400

500

0.00125

0.0025

0.0050

0.0070

0.01

0.0125

10

.0025

.0050

.01

.015

.02

.025

25

.00625

.0125

.025

.0375

.05

.0625

SO

.0125

.025

.05

.075

.1

.125

100

.25

.50

.1

.15

.2

.250

150

.0375

.0750

.15

.225

.3

.375

200

.05

.1

.2

.3

.4

.500

250

.0625

.125

.25

.375

.5

.625

300

.075

.15

.3

.45

.6

.75

500

.125

.25

.5

.75

1.0

1.25

1.5

2.0

2.5

.25
.5
1.0
1000
"See text for correction factors for various types of pump sets."

B. Method of attachment to wheel


In activating any piston pump, it is done ideally, such that straightline motion of the piston rod is
achieved. Any bending in the rod puts undue side loads on the discharge head seal and on the piston
bucket. Straightline motion mechanisms are described and discussed in textbooks, so I will not endeavor
to give details of the common mechanisms.
The books seldom mention however, the practical problems which arise when trying to use such
mechanisms. Nor do they usually compare advantages and disadvantages. I will mention some possible
mechanisms along with the advantages and potential problems.
A slider and crank mechanism (See Fig. 3) is attractive as a simple device with the advantage of requiring
no special techniques to prevent bending moments on the pump plunger. Stroke is easily adjustable by
attaching the crank pin to the wheel shaft via a flange plate with holes drilled at various distances from
the rotation axis, through which the crank pin can be fixed.
Unless a double acting pump is used, the pumping stroke and return stroke will have different forces on
the crank pin resulting in non-uniform wheel rotational speed (unless compensated for by other means such as attaching single acting pumps operating 180 out of phase).
This non-uniform motion can be alleviated to an extent by attaching the slider (pump axis) offset from the
wheel axis. It then becomes a form of quick return mechanism. This, however, increases the side load on
the slider during the return stroke, which necessitates moving the slider bearings apart (increasing the
slider length) to maintain the same slider bearing pressure as with the symmetrical arrangement if bearing
pressure and the resulting frictional drag on the slider become large enough to cause a problem.
Lubrication of the slider bearing presents a problem. Although precautions can limit somewhat the
exposure to water in the bearing, it is unlikely that the bearing can be completely protected. Pressure
grease fittings using a suitably wash-resistant grease might prove suitable.
Packing box style lubrication with oily felt or rags could also be successful. Both methods rely on
periodic attention, which could be of an intolerable frequency. There are also the crank pin and clevis pin
at the slider to be lubricated. Finally, alignment is a potentially tricky problem because of the narrow
tolerance allowable on parallelism of the wheel shaft and crank pin and on perpendicularity of the plane
of the slider crank mechanism with the wheel shaft.

One major advantage compared with the next method discussed is that since the pump body can be fixed
if the alignment is sufficiently accurate, the connection with the distribution pipe can be rigid

TABLE XVI
Quantities of Water Pumped per Stroke for Single Acting Pumps of Various Bore and Stroke Sizes
(Imperial Gallons)
Quantities of Water Pumped per Stroke for Single Acting Pumps of Various Bore and Stroke Sizes
(Imperial Gallons)
Stroke (in.)
Bore (in.)
2 1/4
4
6
8
10
12
.009
.016
.023
.032
.040
.049
1 1/4
.013
.023
.035
.045
.057
.069
1 1/2
.023
.040
.062
.082
.102
.122
2
.035
.064
.095
.127
.159
.191
2 1/2
.052
.092
.139
.184
.230
.278
3
.070
.125
.187
.248
.312
.276
3 1/2
.092
.163
.245
.227
.410
.489
4
.143
.255
.382
.510
.638
.765
5

A second method of attachment is to pivot the pump body about an axis parallel to the wheel shaft (as on
trunnions), attach the pump rod end to the same kind of crank pin as before and let the pump oscillate side
to side as the piston goes up and down. (See Fig. 4). This eases the difficulty of the alignment problem
involving the plane of the crank mechanism previously discussed but introduces new complications.
The pump rod is subjected to side loads. This is ordinarily intolerable at both the gland and the bucket but
fortunately is easily overcome by a simple frame attached to the pump with sliding bearing surrounding
the crank pin which the pump rod end (at the crank pin) then slides in. The bearings absorb all the side
loads required to cause the oscillation, leaving the pump rod loaded linearly only.
Side loads on these slider bearings would be smaller than the side loads on the slider in the slider crank
mounting so that the sliding bearing problems with this technique are somewhat simpler. A serious
objection to this mounting method is the necessity for a flexible connection from the pump to the
distribution pipe.
If the reader intends to build his own pump, which would be likely if considering this particular
arrangement, plan to have the outlet of the pump colinear with the trunnion axis. In this way a simple seal
to allow the pump exit pipe to oscillate in the delivery pipe will suffice. This method of flexible
connection will probably be the most durable.

The scotch yoke mechanism (See Fig. 5) is simple and direct but may require more sophisticated
machining than available equipment will allow. Furthermore, there is the potential danger of excessive
wear and short life if the lubrication is insufficient. This is not generally a suitable mechanism for
unattended use in harsh conditions. A cam activated pump rod is an attractive alternative. It eliminates the
need for any linkage, simplifying the alignment problem and eliminating some parts.

Side loads on a properly designed profile would be very small and a sliding bearing on the outboard end
of the pump rod would easily absorb it. A suitable cam profile is given schemetically in Fig. 6. Force for
the return stroke can easily be supplied by a properly weighted pump rod and the simplest location for
such weight would be immediately above the follower plate. Solid mounting of the pump in this case
allows rigid supply piping to be attached directly to the pump.

A pump bought ready made with a handle can be attached quite simply by a rod suitably aligned between
a crank on the wheel and the free end of the pump handle. Then force and velocity calculations must be
modified.
Various straight line motion linkages are easily constructed. They have the advantage of simplicity and
durability even under harsh working conditions. Many such linkages are discussed in books on Theory of
Machines and Machine Design.
One simple technique to achieve straight line motion seldom seen in texts on machine design is to run a
cable over a pulley such that the end of the cable attached to the pump is colinear with the pump rod. The
other end can be attached to the wheel crank and the cable provides sufficient flexibility that no solid
linkage is needed.
An alternative to this approach is to link the wheel crank to a sector of a pulley sheave in such a way that
the sheave oscillates as the crank rotates. With the cable wrapped far enough around the sector so that the
cable always remains tangent to the sector and fixed there, the free end of the cable can be attached
colinear with the pump rod to provide straight line motion. This is the mechanism used on oil drilling
rigs.
The cable, as a part of the drive mechanism, can be made very long in order to drive pumps located at a
considerable distance from the wheel itself. Such a technique provides the means to power, for instance, a
shallow bore pump in the middle of a village using power generated at a stream some distance away.
C. Piping
For any water distribution system where the water must be transported to a higher elevation, piping is
usually required. There are alternatives such as buckets on an endless belt, etc., but that is outside the
scope of this manual.
The choice will probably fall between polythene and galvanized iron pipe. There are advantages and
disadvantages to both. I shall endeavor to give some helpful information to aid the designer in making the
best choice.
Polythene pipe is available in long (now around 200 metres) lengths so numbers of couplings and joints
are greatly reduced compared to the iron pipe which comes in short lengths (21 1/2 ft typically). It is
flexible (softer, weaker and more elastic in strict engineering terminology) and for this reason is more
susceptable to damage from bush knives, rocks, pig hooves, etc.
Its strength is limited such that it is rated to support at best 300 foot normal working heads at standard
conditions. The strength is strongly temperature dependent however, and at 120F head capacity is down
to 185 ft maximum. It is not fire resistant. Consequently in open country it would probably need to be
buried. If the local soil is very rocky, the burying process must be done with great care to keep the pipe
from suffering rock (penetration) damage. Sand is usually used as a bed and cover.
Iron pipe can generally simply be laid on the ground with rock piles to support it through low spots. It
will support more than 1000 ft heads with plenty of safety margin. For heads to get that high, the system
required will be more sophisticated than can be made by the techniques detailed in this manual.
Prices for the two types are competitive in the higher strength grades of polythene but for low pressure
systems, polythene may be substantially cheaper.
Polythene has a smoother bore so that friction losses are less than with iron pipe, although this would not
likely be a significant factor. It becomes more important in long gravity feed systems.
Weight of a given length is vastly different. 100 ft of high strength 2" polythene weighs 60 lb while 100 ft
2" standard iron pipe weighs 357 lb. Therefore, long distance transport by hand to very remote areas
might influence the decision for polythene even in spite of its other shortcomings.

II. OTHER APPLICATIONS


While water pumping is an obvious use for the water wheel, other machinery can be adapted to use the
mechanical power output of the wheel. It is not the intention of this section to attempt to enumerate all the
possible applications. Rather, I include this section to offset any impression that may have been given by
the preceeding section that water pumping is the most important, or perhaps only use to which the wheel
may be put.
Generation of electricity is a possibility which will probably spring to the minds of most people reading
this manual. There are wheel driven electric generators in operation in Papua New Guinea today but the
number of attempts and failures testify to the fact that it is not a simple, cheap task to make a successful
rig.

The principal difficulties are the speed step-up required for generators and speed regulation. Low voltage
D.C. generation using readily available parts (old auto generators or alternators) avoids the speed
regulation problem. Simple starter-motor-pinion/ flywheel-ring-gear sets could be adequate for speed
step-up at a reasonable cost.
Typical ring gear sets have a lower limit of 10 diametral pitch size teeth which gives a power rating of 10
R.P.M. of about 1/2 h.p. It is, therefore, marginal to expect to produce continuous output from a 12 volt
automobile generator at, say, 60 amps for long periods of time without gear problems.
The small amount of power generated, the need for 12 volt bulbs, resistance losses in long distribution
systems and other problems also mitigate against this being a useful bolt-on accessory. Electricity
generation is better left to the higher speed devices which are more amenable to speed regulation such as
the Banki Turbine of a centrifugal pump being forced to run as a turbine.
Attachment directly to other mechanical machinery can be accomplished by a variety of coupling devices
described in various machine design books. Two circumstances are likely to occur:
1. the machine to be driven will be located some distance from the wheel; and
2. the input shaft of the machine will not easily be aligned with the wheel shaft.
Alignment difficulties are overcome simply and cheaply with old automobile drive shafts and their
attached universal joints. Note that the use of one universal joint will not give constant speed on both
sides. For a constant input speed, the output is alternately faster and slower than the input depending upon
the angle between the two shafts.
The speed variations are small and will generally not be of any consequence. If the speed variations
cannot be tolerated, either a special constant velocity joint (as from the front wheel drive automobile) or
two ordinary U joints must be used, each to compensate for the non-uniform motion of the other.
Flexible shafts are commercially available but are of limited torque carrying capacity.
Solid shafts can transmit torque over considerable distance but require bearings for support and may
therefore be expensive. Virtually any stationary machine which is currently hand-powered could be run
by water wheel power. The means to accomplish the attachment would vary from machine to machine of
course, but only in the case of where the wheel and the machine are separated by long distances should
there be any significant problem.

APPENDIX I
Sample Calculation for Wheel-pump set
The following is an example of the use of this manual to make decisions relating to water wheel for use in
water pumping. The decisions made must be consistent with the bounds placed upon the system by the
village's needs (how much power is required) and the geography and size of the supply stream (how
much power we can expect to get from the wheel).
If the power required is greater than the power that can be generated by the wheel, then the system cannot
work. This example is taken from calculations made for Ilauru village, approximately 15 miles south of
Wau, New Guinea.
One of the possible locations for a wheel is in a stream about 350 feet below the level of the village. The
hill is quite steep and would require about 750 feet of pipe. There is a place in the stream where the water
level drops quite rapidly through a vertical distance of 8 or 10 ft.
The stream is about 10 ft. wide, averages 6 or 9 inches depth and flows about between 1 and 2 ft. per
second (estimated by measuring the time for a leaf to travel a fixed distance). That description establishes
the conditions to determine the maximum wheel size.
The village has about 300 people. Each person now consumes less than 2 gallons of water per day in the
village according to a rough estimate. If the water were pumped into the village, experience in other
countries shows that the consumption would increase. A minimum of 10 gallons per day per person is
sometimes quoted as a minimum viable scheme. Let us calculate for twice that to allow for expansion of
population or of consumption.
1. Total water requirement in gallons per hour 20 gal/person-day x 300 people x day/24 hr = 250 gal/hour
assuming storage facilities at the village to allow larger draw at peak hours.
2. Power required to meet this pumping rate from Table XV. 250 gal/hour at approx. 400 ft. head (350
actual ft. rise + some losses as yet uncalculated) requires about 1/2 h.p. under steady conditions.

3. Depending on the type of pump arrangement used, the wheel will need to be designed for 2 1/2 times
that for a single acting pump, 2 times that for double acting pump or 1 1/2 times that for 2 double acting
pump. Assuming the simplest case of 1 single acting pump we need a wheel of 1 1/4 h.p. potential.
4. Can we get that much power from a water wheel under the stated conditions at the stream? The largest
diameter possible is limited by the drop in the stream in a useable distance -- about 8 feet. An 8 ft. wheel
will operate at about 12 rpm or less (Table VI). The stream has a flow rate of at least
10 ft x 1/2 ft x

1 ft
----sec

5 [ft3]
--------sec

or
5 [ft3] x 6 1/4 gal x 60 sec = 1800 gal
---------------------- -------sec
[ft./3]
min
min

At 1800 gal/min we should be able to produce 2 h.p. at least from an 8 ft. wheel (Table V) or slightly less
depending upon the exact t/r values finally chosen.
Therefore we conclude that the job, in theory, is possible. Had the flow rate been, for example, only 500
gallons per minute, the task of pumping 250 gal per hour to the village would probably have been
impossible.
5. At an estimated 12 rpm and 4 ft. width (maximum usually used is half the diameter) we can estimate
the annulus width necessary (Table II).
1 1/4 h.p. needed
---------------------12 rpm x 4 ft wide
=

0.025 h.p. per rpm per ft of width

In the entry under 8 ft. diameter wheels we see that all annulus widths listed will provide at least that
much power. We now know we can make the wheel less than 4 ft. wide if desired and the annulus width
can be between 3 in. and 12 in.
It is now completely established that an 8 ft. diameter water wheel in this location will do the job
required.
6. If the wheel operates at 12 rpm and the pump is directly coupled so that there is one stroke per rpm
with no added leverage (for instance, as with the wire connection suggested in Part Two, Section IB),
there will be one stroke per revolution.
To accomplish 250 gal/hr we need:
250 gal
----- x
hr

hr
min
------ x ---------60 min
12 strokes

= 0.35 gal/stroke

From Table XVI that means we need 3 1/2 pump with 12" stroke or 4" pump with 9" stroke etc.
7. If we limit the velocity in the pipe to 10 ft/sec then the pipe size with the 3 1/2" pump (chosen because
it is cheaper than the 4" pump) is related to the peak piston velocity and the pump size. From Table XI the
peak piston velocity at 12" stroke 12 rpm is .624 ft/sec. The delivery pipe cross section area must be
approximately.
0.624 x 11

[(3 1/2)2]
1
---------- x --- = Pipe area
4
10

= 0.64 [in2]

This would require a nominal 1" diameter pipe.


8. The pipe would need to be galvanized iron to withstand the pressure of heads exceeding 350 ft. If a
nominal 1" pipe is used, the actual peak velocity is about 7 ft/sec.
The friction head loss would be (Table X)

friction loss = 0.022 x

750
---1/12

[72]
--------- = 150 ft
2 x 32.2

Thus the total peak head causing forces on the pump rod would be 350 (elevation) + 150 (loss) = 500 ft.
Commercial 31/2 m. pumps are fitted with 2 in. pipe outlet holes and if 2 in. pipe is used the loss is much
less because the velocity is less and the diameter is greater.
friction loss = 0.028 x 750
---2/12

[22]
--------- = 8 ft
2 x 32.2

The saving is obviously substantial but the cost of doubling the pipe size may be unattractive.
9. Assuming we use the 1" pipe we find the required pump rod force from Table XII is about 1850 lb. For
a 12" stroke a crank length of 6" is required and so the peak torque on the machine is 925 ft/lb.
From Table 1 we see that this is well within the capacity of the wheel if it is 4 ft. wide.
10. To allow for reasonable future expansion of needs without adding unnecessary weight to the wheel I
would select a 4" annulus. Having done that, the bearing loads are (Table VII) about 500 lb. each.
Assuming the bearings can be located fairly close to the wheel, say 6" away, the solid steel shaft size
required is found from:
[d3] =
16[square root] [(6 x 500)-2] + [(925 x 12)2]
--------------------------------------------------------[pi] (13,000)

d = 1.65 in
Any solid steel shaft larger than this will be satisfactory.

APPENDIX II
An Easily Constructed Piston Pump

by R. Burton
This pump was designed by P. Brown (of the Mechanical Engineering Workshop at the Papua New
Guinea University of Technology) with a view to manufacture in Papua New Guinea. Consequently the
pump can be built up using a minimum of workshop equipment. Most parts are standard pipe fittings
available at any plumbing suppliers.
To avoid having to bore and hone a pump cylinder, a length of copper pipe is used. Provided care is taken
to select an undamaged length and to see that the length is not damaged during construction this system
has proved quite satisfactory.

As can be seen from the cross-sectional diagram, the ends of the pump body consist of copper pipe
reducers silver soldered onto the pump cylinder. This does make disassembly of the pump difficult, but
avoids the use of a lathe.
If a lathe is available, a screwed end could be silver soldered to the upper end of the pump to allow for
simple disassembly.
The piston of the pump consists of a 1/2" thick P.V.C. flange with holes drilled through it (see diagram).
A leather bucket is attached above the piston and together with the holes serves as a non-return valve. In
this type of pump the bucket must be made of fairly soft leather, a commercial leather bucket is not
suitable. Bright steel bar is used as the drive rod and has to be thread cut at its ends using a die.
A galvanized nipple is silver soldered to the top copper reducer of the pump to allow the discharge pipe to
be attached.
An `O' ring seal of the type used to join P.V.C. pipe is used as a seal for the foot valve. This seal does not
require any fixing since it push fits into the lower copper pipe reducer. A 1/2" screwed flange with a plug
in its centre forms the plate for the foot valve. This plate must be restrained from rising up the bore of the
pump by three brass pegs fitted in through the side wall of the pump above the valve plate. These pegs
must be silver soldered in to prevent leakage or movement.
A parts list for a 4" bore x 9" stroke pump is set out below together with a tool list.
Parts
1 only 12" x 4" dia. copper tube
2 only 4" to 1 1/2" copper tube reducers
1 only 1 1/2" galvanized nipple
1 only 1/2" screwed flange
1 only 1/2" plug
1 only 1/2" P.V.C. flange
1 only Rubber `O' ring 4" dia.
1 only piece of 4 1/2" dia. leather
1 only 15" x 1/2" dia. bright steel bar
1/8" dia. brazing rod
Tools
Handi gas kit
Silver solder
Hand drill
1/2" Whitworth die
1/2" Whitworth tap
Hacksaw
Hammer

Overshot Water Wheel


A Design And Construction Manual
I. WHAT IT IS AND WHAT IT IS USED FOR
BACKGROUND
Improved use of water as a power source has potential for much of the developing world. There are few
places where water is not available in quantities sufficient for power generation. Almost any flowing
water--river, brook, or outlet of a lake or pona--can be put to work and will provide a steady source of
energy. Fluctuations in the rate of flow usually are not too large and are spread out over time; water flow
is far less subject to quick changes in energy potential and is available 24 hours a day.
The uses of energy from water are about the same as those for energy from the wind--electrical
generation and mechanical power. Water-powered turbines attached to generators are used to generate
electricity; waterwheels are generally used to power mechanical devices such as saws and machines for
grinding grain.

Development of water power can be advantageous in communities where the cost of fossil fuels is high
and access to electric transmission lines is limited.

POSSIBLE APPLICATIONS
The cost of employing water power can be high. As with any energy project, you must consider carefully
all options. The
potential for power generation of the water source must be carefully matched with what it will power. For
example, if a windmill and a waterwheel can be constructed to fill the same end use, the windmill may
well require less time and money. On the other hand, it may be less reliable.
Using water power requires:
1) a constant and steady flow of water, and
2) sufficient "head" to run the waterwheel or turbine, if such is being used.
"Head" is the distance the water falls before hitting the machine, be it waterwheel, turbine, or whatever. A
higher head means more potential energy.
There is a greater amount of potential energy in a larger volume of water than in a smaller volume of
water. The concepts of head and flow are important: some applications require a high head and less flow;
some require a low head but a greater flow.
Many water power projects require building a dam to ensure both constant flow and sufficient head. It is
not necessary to be an engineer to build a dam. There are many types of dams, some quite easy to build.
But any dam causes changes in the stream and its surroundings, so it is best to consult someone having
appropriate expertise in construction technique.
It is important to keep in mind that there can be substantial variation in the available flow of water, even
with a dam to store the water. This is especially true in areas with seasonal rainfall and cyclical dry
periods. Fortunately, in most areas these patterns are familiar.
Waterwheels have particularly high potential in areas where fluctuations in water flow are large and
speed regulation is not practical. In such situations, waterwheels can be used to drive machinery which
can take large fluctuations in rotation and speed. Waterwheels operate between 2 and 12 revolutions per
minute and usually require gearing and belting (with related friction loss) to run most machines. (They
are most useful for slow-speed) applications, e.g., flour mills, agricultural machinery, and some pumping
operations.
A waterwheel, because of its rugged design, requires less care than a water turbine. It is self-cleaning, and
therefore does not need to be protected from debris (leaves, grass, and stones).
Capital and labour costs can vary greatly with the way the power is used. For example, an undershot
waterwheel in a small stream can be fairly easy and inexpensive to build. On the other hand, the set-up
for generating electricity with a turbine can be complicated and costly. However, once a water power
device is built and in operation, maintenance is simple and low in cost: it consists mainly of lubricating
the machinery and keeping the dam in good condition. A well built and well situated water power
installation can be expected to last for 20-25 years, given good maintenance and barring major
catastrophes. This long life is certainly a factor to be figured into any cost calculation.

II. DECISION FACTORS


Applications:
water pumping.
Low-speed machinery applications such as grist mills, oil presses, grinding machines, coffee
hullers, threshers, water pumps, sugar cane presses, etc.
Advantages:
Can work over a range of water flow and head conditions.
Very simple to build and operate.
Virtually no maintenance required.
Considerations:
Not advisable for electrical generation or high-speed machinery applications.
For optimum life expectancy water-resistant paints are needed.
COST ESTIMATE(*)
$100 to $300 (US, 1979) including materials and labour.
(*) Cost estimates serve only as a guide and will vary from country to country.

III. MAKING THE DECISION AND FOLLOWING THROUGH


When determining whether a project is worth the time, effort, and expense involved, consider social,
cultural, and environmental factors as well as economic ones. What is the purpose of the effort? Who will
benefit most? What will the consequences be if the effort is successful? And if it fails?
Having made an informed technology choice, it is important to keep good records. It is helpful from the
beginning to keep data on needs, site selection, resource availability, construction progress, labour and
materials costs, test findings, etc. The information may prove an important reference if existing plans and
methods need to be altered. It can be helpful in pinpointing "what went wrong?" And, of course, it is
important to share data with other people.
The technologies presented in this and the other manuals in the energy series have been tested carefully,
and are actually used in many parts of the world. However, extensive and controlled field tests have not
been conducted for many of them, even some of the most common ones. Even though we know that these
technologies work well in some situations, it is important to gather specific information on why they
perform properly in one place and not in another.
Well documented models of field activities provide important information for the development worker. It
is obviously important for a development worker in Colombia to have the technical design for a machine
built and used in Senegal. But it is even more important to have a full narrative about the machine that
provides details on materials, labour, design changes, and so forth. This model can provide a useful frame
of reference.
A reliable bank of such field information is now growing. It exists to help spread the word about these
and other technologies, lessening the dependence of the developing world on expensive and finite energy
resources.
A practical record keeping format can be found in Appendix VI.

IV. PRE-CONSTRUCTION CONSIDERATIONS


The two most common types of waterwheels are the undershot and the overshot versions.

UNDERSHOT WATERWHEEL
The undershot waterwheel (see Figure 1) should be used with a head of 1.5 to 10 ft and flow rates from
10 to 100 cu ft per second. Wheel diameter should be three to four times the head and is usually between
6 and 20 ft.
Rotational speeds of the wheel are from 2-12 revolutions per minute; smaller wheels produce higher
speeds. The wheel dips from 1-3 ft into the water.
Efficiency is in the range of 60-75 percent.

OVERSHOT WATERWHEEL

The overshot waterwheel (see Figure 2) is used with heads of 10-30 ft and flow rates from 1-30 cu ft per
second.
Water is guided to the wheel through a wood or metal flume. A gate at the end of the flume controls the
water flow to the wheel.

Wheel width can be fixed depending upon the amount of water available and the output needed. In
addition, the width of the waterwheel must exceed the width of the flume by about 15cm (6") because the
water expands as it leaves the flume.
The efficiency of a well constructed overshot waterwheel can be 60-80 percent.
Overshot wheels are simple to construct, but they are large and they require a lot of time and material--as
well as a sizeable workspace. Before beginning construction, it is a good idea to be sure facilities are or
will be available for transporting the wheel and lifting it into place.
Even though an overshot wheel is simple to construct and does not require extreme care in cutting and
fitting, it must be strong and sturdy. Its size alone makes it heavy, and in addition to its own weight, a
wheel must support the weight of the water. The high torque delivered by the wheel requires a strong
axle--a wooden beam or (depending on the size of the wheel) a car or tractor axle. Attention to these
points will help prevent problems with maintenance.
Large waterwheels may be made much like a wagon wheel--with a rim attached to spokes. A smaller
wheel may be made of a solid disc of wood or steel. Construction of a wheel involves the assembly of
four basic parts: the disc or spokes of the wheel itself, the shrouds or sides of the buckets that hold the
water, the buckets, and the mounting framework. Other parts are determined by the work the wheel is
intended to do and may include a drive for a pump or grinding stone or a system of gears and pulleys for
generating electricity.
Before a wheel is constructed, careful consideration should be given to the site of the wheel and the
amount of water available. Because overshot wheels work by gravity, a relatively small flow of water is
all that is needed for operation. Even so, this small flow must be directed into a flume or chute. Doing this
often requires construction of a small dam.
The overshot waterwheel derives its name from the manner in which it is activated by the water. From a
flume mounted above the wheel, water pours into buckets attached to the edge of the wheel and is
discharged at the bottom. An overshot wheel operates by gravity: the water-filled buckets on the
downward side of the wheel over-balance the empty buckets on the opposite side and keep the wheel
moving slowly.
In general, overshot waterwheels are relatively efficient mechanically and are easily maintained. Their
slow speed and high torque make them a good choice to operate such machinery as grist mills, coffee
hullers, and certain water pumps. They may even be used for generating small amounts of electricity.
Electrical generators require a series of speed multiplying devices that also multiply the problems of cost,
construction, and maintenance.
Such a wheel should be located near, but not in, a stream or river. If a site on dry ground is chosen, the
foundation may be constructed dry and the water led to the wheel and a tailrace excavated (see Figure 3).

Efficiency of the wheel depends on efficient and practical design considerations. The wheel must use the
weight of the water through as much of the head as possible. The buckets should not spill or sling water
until very near the tailwater.

The experience of the people at an isolated hospital in rural Malawi serves to illustrate many of the
questions, both technical and cultural, that go into the development of a water power unit.
A failed cassava crop in the area led to the substitution of a new dietary staple--corn (maize). But the
nearest mill for grinding the corn was a 49-kilometer (30-mile) walk away. Clearly something needed to
be done to make milling facilities more accessible to the people.
A diesel-powered mill was too expensive and too difficult to maintain in that remote region. The river
flowing past the hospital seemed to hold the promise of a power source, but, again, commercial water
turbines proved too costly. Some kind of waterwheel seemed to provide an appropriate choice.
Development of the water power site involved the combined efforts of VITA and five VITA Volunteers,
a missionary engineer in another area of Malawi, and OXFAM, another international development
agency. Some data was also supplied by commercial milling firms. Much of the labour was volunteered
by local people.
Correspondence between and among the participants involved choice of type of wheel, determining how
to provide enough head to develop enough power to do the job, construction of the wheel, and selecting
the proper burrs or grindstones.
Both VITA and OXFAM strongly recommended an overshot wheel for the reasons cited earlier: ease of
construction and maintenance, reliability, and mechanical efficiency. With this comparison as a guide, the
overshot wheel was chosen.
Power to run the grain mill required a head of about 427cm (14 ft, which would accommodate a wheel
nearly 361cm (12 ft) across. The higher head necessary for the overshot wheel made it necessary to clear
away additional boulders from the river bed, but this original investment in labour was more than
returned by the increased efficiency of the wheel.
Additional correspondence (except for a couple of visits by the missionary engineer, the entire problemsolving process was handled by mail!) determined the precise shape, angle, size, and numbers of the

buckets on the wheel. Also necessary was the design of a system of pulleys to transfer the power of the
wheel to the milling operation.
As the wheel was constructed, attention was given to the grindstones. Granite found in the area seemed
ideal, but proved to be too difficult for local stone cutters to deal with and yet not durable enough to be
worth the time. Advice was sought from a millwright in New York and a variety of commercial milling
firms. Ultimately a small commercial mill was chosen, with continued study going into preparing
traditional stones.
In one of the last letters, hospital staff related that the wheel and mill were in place and operating. And
from experience gained in this project they were already considering the possibility of constructing
turbines to generate electricity.

SITE SELECTION
A careful analysis of the proposed site of the waterwheel is an important early step before construction
begins. Whether it is a good idea to try to harness a stream depends on the reliability and quantity of the
flow of water, the purpose for which power is desired, and the costs involved in the effort.
It is necessary to look at all factors carefully. Does the stream flow all year round--even during dry
seasons? How much water is available at the driest times? What will the power do--grind grain, generate
electricity, pump water? These questions and others must be asked.
If a stream does not include a natural waterfall of sufficient height, a dam will have to be built to create
the 'head' necessary to run the wheel. Head is the vertical distance which the water falls.
The site of the dam and wheel will affect the amount of head available. Water power can be very
economical when a dam can be built into a small river with a relatively short (less than 100 ft) conduit
(penstock for conducting water to the waterwheel).
Development costs can be fairly high when such a dam and pipeline can provide a head of only 305cm
(10 ft) or less. While a dam is not required if there is enough water to cover the intake of a pipe or
channel at the head of the stream where the dam would be placed, a dam is often necessary to direct the
water into the channel intake or to get a higher head than the stream naturally affords.
This, of course, increases expense and time and serves as a very strong factor in determining the
suitability of one site over another.
A thorough site analysis should include collection of the following data:
Minimum flow in cubic feet or cubic meters per second.
Maximum flow to be utilized.
Available head in feet or meters.
Site sketch with elevations, or topographic map with site sketched in.
Water condition, whether clear, muddy, sandy, etc.
Soil condition, the velocity of the water and the size of the ditch or channel for carrying it to the
works depends on soil condition.
Measurements of stream flow should be taken during the season of lowest flow to guarantee full power at
all times. Some investigation of the stream's history should be made to determine if there are perhaps
regular cycles of drought during which the stream may dry up to the point of being unusable.
Appendices I and II of this manual contain detailed instructions for measuring flow, head, etc., and for
building penstocks and dams. Consult these sections carefully for complete directions.

POWER OUTPUT
The amount of water available from the water source can be determined to assist in making the decision
whether to build. Power may be expressed in terms of horsepower or kilowatts. One horsepower is equal
to 0.7455 kilowatts; one kilowatt is about one and a third horsepower.
The gross power, or full amount available from the water, is equal to the useful power plus the losses
inherent in any power scheme. It is usually safe to assume that the net or useful power in small power
installations will only be half of the available gross power due to water transmission losses and the
gearing necessary to operate machinery.
* Gross power is determined by the following formula:
In English units:
Gross Power (horsepower) =

Minimum Water Flow (cu ft/sec) X Gross Head (ft)


-------------------------------------------------------8.8
In Metric units:
Gross Power (metric horsepower) =
1,000 Flow (cu m/sec)
----------------------- X Head (m)
75
* Net power available at the turbine shaft is:
In English units:
Net Power =
Minimum Water Flow X Net Head(*)
----------------------------------------- X Turbine Efficiency
8.8
In Metric Units:
Net Power =
Minimum Water Flow X Net Head(*)
----------------------------------------- X Turbine Efficiency
75/1,000
(*) The net head is obtained by deducting the energy losses from the gross head. These losses are
discussed in Appendix I. When it is not known, a good assumption for waterwheel efficiency is 60 percent.

APPLICATIONS
While water pumping is an obvious use for the waterwheel, other machinery can be adapted to use the
mechanical power output of the wheel. Almost any stationary machine which is currently hand-powered
could be run by waterwheel power. Only in the case where the wheel and the machine are separated by
long distances should there be any significant problem.
One problem which can occur when the machine is located some distance from the wheel is that the drive
shaft of the machine will not easily be aligned with the waterwheel shaft. Alignment difficulties can be
overcome simply and cheaply with old automobile rear axle assemblies, with the gears welded or jammed
to give constant speed on both sides.
If the water supply to the wheel fluctuates, the speed of the wheel will vary. These speed variations are
small and will generally not be of any consequence. If the variable speeds create problems, either a
special constant velocity joint (as from the front wheel drive automobile) or two ordinary U joints must
be used, each to compensate for the different motion of the other.
Flexible shafts are commercially available but are of limited torque-carrying capacity.
Solid shafts can transmit torque over considerable distance but require bearings for support and are
expensive.
Generation of electricity is a possibility which will probably spring to the minds of most people reading
this manual. There are waterwheel-driven electric generators in operation today, but the number of failed
attempts testifies to the fact that it is not a simple, inexpensive project.

RECORDS
The need for power should be documented, and the measurements taken for the site analysis should be
recorded. Costs of construction and operation can be compared to the benefit gained from the device to
determine its real worth. (In making comparisons, don't forget to include the pond or lake created by the
dam--it can be used to water livestock, raise fish, or irrigate fields.)

MATERIALS AND TOOLS


A simple, relatively economical 112cm (5 ft) wheel for pumping water can be made out of a disc of
heavy plywood to which the buckets and shrouds are attached. Plywood is chosen because it is easy to
use and relatively accessible; however, it does require special treatment to avoid deterioration and, in
some places, may be quite expensive. The shaft of the wheel can be made either from metal or wood: the
rear axle from an automobile may be used but, in most cases, axles are available only at great expense.
Lumber for shrouds, buckets, and rim reinforcement may be of almost any type available; hardwood is
preferable. Ordinary wood saws, drills, and hammer are used in construction. Welding equipment is
convenient if an automobile rear axle is being used. Materials for the dam and mounting structure should

be chosen from whatever is at hand, based on the guidelines in this manual. While materials for the wheel
may vary with what is available, they should include:
Materials
2cm thick plywood(*)--at least 112cm square.
6mm thick plywood(*)--122cm X 244cm sheet.
703cm total length of 3cm X 6cm boards to reinforce the edge of the disc.
703cm total length of 2cm X 30cm boards for the shrouds.
438cm total length of 2cm X 30cm boards for buckets.
703cm total length of 6mm X 20cm plywood* to reinforce the outside of the shrouds.
110cm long 5cm dia solid steel shaft or 9cm sq hardwood shaft. (Automobile rear axle is
optional.)
5cm dia steel hubs (2) for steel shaft.
10 litres asphalt patching compound (or tar).
Timbers and lumber for support structure as needed, nails, tin cans, bolts.
(*) Marine-grade plywood is preferred; waterproofed exterior-grade can be used.
Tools
Protractor
Wood saw
Wood drill/bits
Hammer
Welding equipment (optional)

V. CONSTRUCTION
PREPARE THE DIAMETER SECTION
* Make a disc out of the 2cm thick plywood 112cm in dia. This is done using the meter stick.
* Nail one end of the ruler to the center of the plywood sheet.
* Measure 56cm from the nail and attach a pencil to the ruler.
* Scribe a circle and cut out disc with a wood saw (see Figure 4 below).

* Divide the circle in half and then into quarters using a pencil and straight edge.
* Divide each quarter into thirds (30[degrees] intervals on protractor). The finished disc should look like
Figure 5. The twelve reference lines will be used to guide the positioning of the buckets.

* Take 25-40cm lengths of 2cm X 3cm X 6cm lumber and nail them around the outside diameter of the
wood disc on both sides so that the outer edge projects slightly beyond the rim of the disc (see Figure 6).

* Cut the 6mm thick X 122cm X 244cm plywood sheets into six strips 40.6cm wide X 122cm long.
* Bend and nail three of the strips around the disc so that they overhang equally on both sides.
* Bend and nail a second layer over the first, staggering the joints so as to give added strength and
tightness (see Figure 7). These layers form what is called the sole plate or backside of the buckets which
will be attached later.

PREPARE THE SHROUDS


* Cut the shrouds, or sides, of the buckets from 2cm X 30cm wide boards. Nail one end of the meter stick
to a piece of lumber. Measure 57.2cm from this nail. Drill 6mm hole and attach a pencil.
* Measure 20.5cm from this pencil, drill 6mm hole and attach another pencil. This becomes a compass
for making the shrouds (see Figure 8).

* Take 2cm X 30cm boards and scribe the outline of the shroud on, the wood. Cut out enough of the
shrouds to fit around both sides of the disc. Shroud edges will have to be planed to fit.

* Nail the shroud pieces flush to the edge of the sole plate from the back side of the sole plate.
* Use the "compass" trace and cut out a second set of shrouds, or shroud covers, from 6mm thick
plywood.
* Nail the plywood shroud covers on the outside of the first shrouds, with the joints overlapped (see
Figure 9). Be sure that the bottom edge of this second set of shrouds is flush with the bottom edge of the
first layer of the sole plate.

* Fill in any cracks and seams with the asphalt patching compound or waterproof sealer. The finished
wheel will look something like a cable spool (see Figure 10).

PREPARE THE BUCKETS

* Make the front sides of the twelve (12) buckets from hardwood boards 2cm X 30cm. The width of the
front board will be 36.5cm.
* Make the bottom sections of the buckets from hardwood boards 2cm X 8cm. The length of each board
will be 36.5cm.
* Cut the bottom of each 30cm section at a 24[degrees] angle from the horizontal and the top edge at a
45[degrees] angle from the horizontal as shown in Figure 11 before putting the two sections together.

* Nail the buckets together (see Figure 12). Each bucket should have an inside angle of 114 . Place each
bucket between the shrouds. Using the reference lines scribed on the disc earlier, match one bucket to
each line as shown in Figure 13. The buckets can then be nailed in place.

* Fill in all cracks with the asphalt patching compound.

MAKE THE WOOD BEARINGS


Bearings, for attaching the shaft to the wheel, will last longer if they are made from the hardest wood
available locally. Generally, hardwoods are heavy and difficult to work. A local wood-craftsman should
be able to provide information on the hardest woods. If there is doubt concerning the hardness or the selflubricating quality of the wood that is going to be used in the bearings, thoroughly soaking the wood with
oil will give longer life to the bearings.
Some advantages in using oil-soaked bearings are that they:
Can be made from locally available materials.
Can be made by local people with wood-working skills.
Are easily assembled.
Do not require further lubrication or maintenance in most cases.
Are easily inspected and adjusted for wear.
Can be repaired or replaced.
Can provide a temporary solution to the repair of a more sophisticated production bearing.
The oiliness of the wood is important if the bearing is not going to be lubricated. Woods having good
self-lubricating properties often are those which:
Are easily polished.
Do not react with acids (e.g., teak).
Are difficult to impregnate with preservatives.
Cannot be glued easily.
Usually the hardest wood is found in the main trunk just below the first branch. Wood freshly cut should
be allowed to dry for two to three months to reduce moisture content. High moisture content will result in
a reduction in hardness and will cause greater wear.
SIZE OF THE BEARING
The length of the wood bearings should be at least twice the shaft diameter. For example, for the 5cm dia
axle or shaft of the waterwheel presented here, the bearing should be at least 10cm long. The thickness of

the bearing material at any point should be at least the shaft diameter (i.e., for a 5cm dia shaft a block of
wood 15cm X 15cm X 10cm long should be used).
Split block bearings (see Figure 14) should be used for the waterwheel because it is a heavy piece of
equipment and can cause a great deal of wear. These bearings are simple to make and replace.

The following steps outline the construction of a split-block bearing:


* Saw timber into an oblong block slightly larger than the finished bearing to allow for shrinkage.
* Bore a hole through the wood block the size of the axle/ shaft diameter.
* Cut block in half and clamp firmly together for drilling.
* Drill four 13mm or larger holes for attaching bearing to bearing foundation. After drilling, the two
halves should be tied together to keep them in pairs.
* Impregnate the blocks with oil.
* Use an old 20-litre (5-gal) drum filled two-thirds full with used engine oil or vegetable oil.
* Place wood blocks in oil and keep them submerged by placing a brick on top (see Figure 15).

* Heat the oil until the moisture in the wood is turned into steam--this will give the oil an appearance of
boiling rapidly.
* Maintain the heat until there are only single streams of small pin-sized bubbles rising to the oil's surface
(see Figure 16).

This may take 30 minutes to 2 hours, or longer, depending on the moisture content of the wood.
Soon after heating the bearing blocks in oil, many surface bubbles one-inch in diameter, made from a
multitude of smaller bubbles, will appear on the surface.
As the moisture content of blocks is reduced, the surface bubbles will become smaller in size.
When the surface bubbles are formed from single streams of pin-sized bubbles, stop heating.
* Remove the heat source and leave the blocks in the oil to cool overnight. During this time the wood will
absorb the oil.
BE VERY CAREFUL IN HANDLING THE CONTAINER OF HOT OIL.
* Remove wood blocks from the oil, reclamp and rebore the holes as necessary to compensate for
shrinkage that may have taken place. The bearings are now ready to be used.

(Calculations for shaft and bearing sizes for larger waterwheels are provided in Appendix IV.)

ATTACH METAL OR WOOD SHAFT TO WHEEL


Metal Shaft
* Drill or cut out a 5cm dia round hole in the center of the wheel.
* Attach 5cm dia steel hubs as shown in Figure 17 using four 20mm X 15cm long bolts.

* Insert 110cm long metal shaft through the wheel center so that the shaft extends 30cm from one edge of
the shroud and 38.2cm from the other edge (see Figure 18).

* Weld the shaft to the hub assembly on both sides as shown in Figure 19.

Wood Shaft
* Drill and carefully cut out a 9cm square hole in the center of the wheel.
* Measure 49cm from one end of the 110-cm long wood shaft and mark with a pencil. Measure 59cm
from other end of the shaft and do the same. Turn the shaft over and repeat the procedure. There should
be 2cm between the two marks.
* Cut grooves 3cm wide X 1cm deep on both sides of the shaft as shown below in Figure 20.

* Cut the 9cm shaft to 5cm dia only at the bearing (see Figure 21).

This step will take some time. A tin can 5cm in diam or the bearing itself can be used to gauge the cutting
process. The finished shaft must be sanded and made as round and smooth as possible to prevent
excessive or premature wear on the bearing.
* Insert wood shaft through wheel center so that the grooves show on either side of the wheel disc.
* Fit 3cm X 6cm X 15cm boards into the grooves so that they fit tightly. Tack each board to disc using
nails to ensure a tight fit in the groove.
* Drill two 20mm dia holes through each 3cm X 6cm boards and disc. Insert 20mm dia X 10cm long
bolts with washer through the disc and attach with washer and nut (see Figure 22 and Figure 23).
Remove nails.

* The wheel is now ready to be mounted.

CONSTRUCTING MOUNTINGS AND TAILRACE


Stone or concrete pillars make the best mounting for the waterwheel. Heavy wood pilings or timber also
have been used successfully. The primary determinant is, of course, local availability. Foundations should
rest on a solid base--firm gravel or bedrock if possible to avoid settling. Large area footings will also
help, and will prevent damage from stream erosion. If one end of the shaft is supported at the power plant
building, this support should be as solid as the outer pillar.
Provision should be made for periodic adjustment in the alignment of the bearings in case one of the
supports should settle or slide. Wood blocks can be used to mount the bearings, and these can be changed
to adjust for any differences in elevation or placement. It is important that bearings and wheel shift be
kept in perfect alignment at all times.
If the discharge or tailwater is not immediately removed from the vicinity of the wheel, the water will
tend to back up on the wheel causing a serious loss of power. However, the drop necessary to remove this
water should be kept at a minimum in order to lose as little as possible of the total available head.
The distance between the bottom of the wheel and the tailrace should be 20-30cm (4-6"). The tailrace or
discharge channel should be smooth and evenly shaped down the stream bed below the wheel (see Figure
24).

MOUNTING THE WHEEL


Attach the bearings to the shaft and lift the wheel onto mounting pillars. Align the wheel vertically and
horizontally through the use of wood blocks under the bearings. Once alignment has been done, drill
through four holes in the bearing into the wood shim and mounting pillar. Attach the bearings to the
pillars using lag/anchor bolts in the case of concrete pillars or lag/anchor screws 13mm dia X 20cm long
if wood pilings are used.
In mounting the shaft in the bearings, carefully avoid damage to the bearings and shaft. The shaft and
bearings must be accurately aligned and solidly secured in place before the chute is assembled and
located.
The wheel must be balanced in order to run smoothly, without uneven wear, or excess strain on the
supports. When the wheel is secured on the mountings, it should turn easily and come to a smooth, even
stop. If it is unbalanced, it will swing back and forth for a time before stopping. If this should occur, add a
small weight (i.e., several nails or a bolt), at the top of the wheel when it is stopped. With care, enough
weight can be added to balance the wheel perfectly.

MOUNTING THE WHEEL--VEHICLE AXLE (Optional)


Take a rear axle from a full-sized car and fix the differential gears so the two axles turn as one unit. You
can jam these gears by welding so they don't operate. Cut off one axle and the axle housing to get rid of
the brake assembly if you wish.
The other axle should be cleaned of brake parts to expose the hub and flange. You may have to knock the
bolts out and get rid of the brake drum. The wooden disc of the waterwheel needs to have a hole made in

its center to fit the car wheel hub closely. Also it should be drilled to match the old bolt holes and bolts
installed with washers under the nuts.
Before mounting the wheel in place, have a base plate welded to the axle housing (see Figure 25). It
should be on what is to be the underside, with two holes for 13mm lag screws. Make some kind of anchor
to hold the opposite housing in place.

WATER DELIVERY TO WHEEL


For highest efficiency, water must be delivered to the wheel from a chute placed as close to the wheel as
possible, and arranged so that the water falls into the buckets just after they reach upper dead center (see
Figure 26). The relative speed of the buckets and the water are very important.

The speed of the wheel will be reduced as the load it is driving increases. When large changes take place
in the load, it is necessary to change the amount of water or the velocity of its approach to the wheel. This
is done by a control gate located near the wheel, which can be raised or lowered easily and fixed at any
position to give moderately accurate adjustment.
The delivery chute should run directly from the control gate to the waterwheel, and be as short as
construction will permit (30cm-91cm long is best). A little slope is necessary to maintain the water
velocity (1% or 30cm in every 3000cm will be satisfactory).
Flat-bottomed chutes are preferable. Even when water is delivered through a pipe, this should terminate
in a control box and delivery made to the wheel through an open, flat-bottomed chute (see Figure 27).
The tip of the chute should be perfectly straight and level, and lined with sheet metal to prevent wear.

The chute should not be as wide as the waterwheel. This allows air to escape at the ends of the wheel as
water enters the buckets. The width of the chute is usually 10-15cm (4-6") narrower than the width of the
wheel. (In this case where the bucket width is 36.5cm the chute width will be 22-26cm.)

MAINTENANCE
All plywood parts must be waterproofed to keep them from deteriorating. Other wood parts may be
painted or varnished for a protective coating. This helps extend the life of the wheel. wheel. Periodic
repainting may be needed. Except for the plywood portions, the decision to paint can be made on purely
economic grounds. If a very durable wood has been used initially, painting is a luxury. If a somewhat less
durable species is used, painting is probably cheaper and easier than early replacement or repair of the
wheel.
The only major maintenance problem is in bearing wear. Generous allowances have been made in bearing
size but the bearings will still wear. When worn, the two halves can be interchanged; after further wear,
the life of the bearing can be extended by planing off a small amount of wood from the matching faces.
This will drop the wheel from its original position. Inserting wood or shimming under the bearing block
with metal plates will compensate for this. Bearing replacement, when the block is completely worn
through, is a simple matter.
Generally speaking, the bearing should be lubricated as needed. Oils/grease/vegetable oils applied
periodically in small amounts will slow the wear rate.

VI. GLOSSARY
CATASTROPHE--A great and sudden disaster of calamity.
CYCLICAL DRY PERIODS--A periodically repeated sequence of environmental conditions where there
is a lack of rainfall or water supply.
DIA (DIAMETER)--A straight line passing through the center of a circle and meeting at each end of the
circumference.
EMBED--To fix firmly in a surrounding mass. Head--Measurement of the difference in depth of a liquid
at two given points (see Appendix I). FLUME--A channel or chute for directing the flow of water.
FLUCTUATIONS--Irregular variations or instability of a regular process.

GRAVITY--The force of attraction that causes terrestial bodies to tend to fall toward the center of the
earth.
RACK AND PINION--A device to convert rotary motion to linear motion.

SHROUD--A device that covers, conceals, or protects something.


SLUICE--A man-made water channel with a valve or gate to regulate the flow.
SPROCKETS--Any of various toothlike projections arranged on a wheel rim to engage the links of a
chain.
TELESCOPIC--Capable of being made longer or shorter by the sliding of overlapping tubular sections.
TOPOGRAPHIC MAP--A map showing the configuration of a place or region, usually by the use of
contour lines.

VII. FURTHER INFORMATION RESOURCES


Cloudburst Press Ltd. Cloudburst manual, 1973. Cloudburst Press Ltd., Mayne Island, British Columbia,
VON 2JO Canada.
This manual, written by "homesteaders" in the Pacific Norethwest, has about 30 pages dealing with
various aspects of water power. It covers measuring potential power, dams, and designs and construction
of waterwheels. Highly readable and eminently practical, it is written by and for "do-it-yourselfers"
working with limited resources. Also has excellent illustrations.
Hamm, Hans W. Low Cost Develoment of Small Water Power Sites, 1967. VITA, 3706 Rhode Island
Avenue, Mount Rainier, Maryland 20822 USA.
Written expressly to be used in developing areas, this manual contains basic information on measuring
water power potential, building small dams, different types of turbines and waterwheels, and several
necessary matehmatical tables. Also has some information on manufactured turbines available. A very
useful book.
Monson, O.W., and Hill, Armin J. Overshot and Current Water Wheels, January 1942. Bulletin 398,
Montana State College Agricultural and Experimental Station, Bozeman, Montana, USA.
Written for the use of farmers and ranchers, this bulletin tells how to build "homemade" waterwheels
from wood and scrap metal, as the emphasis is on simplicity and low cost. A good guide for building and
installing overshot and undershot waterwheels, it is profusely illustrated and contains many practical hints
for consideration.
Ovens, William G. A Design Manual for Waterwheels, 1975. VITA, 3706 Rhode Island Avenue, Mount
Rainier, Maryland 20822 USA.
The basic manual for waterwheel design and construction. Includes both theoretical and practical
considerations, and is written to be used by people with a limited technical understanding. Also has a
section on waterwheel applications as well as 16 highly useful tables and several schematic diagrams.

VIII. CONVERSION TABLES


UNITS OF LENGTH
1 Mile = 1760 Yards = 5280
Feet 1 Kilometer = 1000 Meters = 0.6214 Mile
1 Mile = 1.607 Kilometers 1 Foot = 0.3048 Meter
1 Meter = 3.2808 Feet = 39.37 Inches
1 Inch = 2.54 Centimeters
1 Centimeter = 0.3937 Inches
UNITS OF AREA
1 Square Mile = 640 Acres = 2.5899 Square Kilometers
1 Square Kilometer = 1,000,000 Square Meters = 0.3861 Square Mile
1 Acre = 43,560 Square Feet
1 Square Foot = 144 Square Inches = 0.0929 Square Meter

1 Square Inch = 6.452 Square Centimeters


1 Square Meter = 10.764 Square Feet
1 Square Centimeter = 0.155 Square Inch
UNITS OF VOLUME
1.0 Cubic Foot = 1728 Cubic Inches = 7.48 US Gallons
1.0 British Imperial Gallon = 1.2 US Gallons
1.0 Cubic Meter = 35.314 Cubic Feet = 264.2 US Gallons
1.0 Liter = 1000 Cubic Centimeters = 0.2642 US Gallons
1.0 Metric Ton = 1000 Kilograms = 2204.6 Pounds
1.0 Kilogram = 1000 Grams = 2.2046 Pounds
1.0 Short Ton = 2000 Pounds
UNITS OF PRESSURE
1.0 Pound per square inch = 144 Pound per square foot
1.0 Pound per square inch = 27.7 Inches of water*
1.0 Pound per square inch = 2.31 Feet of water*
1.0 Pound per square inch = 2.042 Inches of mercury*
1.0 Atmosphere = 14.7 Pounds per square inch (PSI)
1.0 Atmosphere = 33.95 Feet of water*
1.0 Foot of water = 0.433 PSI = 62.355 Pounds per square foot
1.0 Kilogram per square centimeter = 14.223 Pounds per square inch
1.0 Pound per square inch = 0.0703 Kilogram per square centimeter
(*) At 62 degrees Fahrenheit (16.6 degrees Celsius).
UNITS OF POWER
1.0 Horsepower (English) = 746 Watt = 0.746 Kilowatt (KW)
1.0 Horsepower (English) = 550 Foot pounds per second
1.0 Horsepower (English) = 33,000 Foot pounds per minute
1.0 Kilowatt (KW) = 1000 Watt = 1.34 Horsepoer (HP) English
1.0 Horsepower (English) = 1.0139 Metric horsepower (cheval-vapeur)
1.0 Metric horsepower = 75 Meter X Kilogram/Second
1.0 Metric horsepower = 0.736 Kilowatt = 736 Watt

APPENDIX I
SITE ANALYSIS
This Appendix provides a guide to making the necessary calculations for a detailed site analysis.
Data Sheet
Measuring Gross Head
Measuring Flow
Measuring Head Losses

DATA SHEET
1. Minimum flow of water available in cubic feet per second (or cubic meters per second).
2. Maximum flow of water available in cubic feet per second (or cubic meters per second).
3. Head or fall of water in feet (or meters).
4. Length of pipe line in feet (or meters) needed to get the required head.
5. Describe water condition (clear, muddy, sandy, acid).
6. Describe soil condition (see Table 2).
7. Minimum tailwater elevation in feet (or meters).
8. Approximate area of pond above dam in acres (or square kilometers).
9. Approximate depth of the pond in feet (or meters).
10. Distance from power plant to where electricity will be used in feet (or meters).
11. Approximate distance from dam to power plant.
12. Minimum air temperature.
13. Maximum air temperature.
14. Estimate power to be used.

15. ATTACH SITE SKETCH WITH ELEVATIONS, OR TOPOGRAPHIC MAP WITH SITE
SKETCHED IN.
The following questions cover information which, although not necessary in starting to plan a water
power site, will usually be needed later. If it can possibly be given early in the project, this will save time
later. 1. Give the type, power and speed of the machinery to be driven and indicate whether direct, belt, or
gear drive is desired or acceptable.
2. For electric current, indicate whether direct current is acceptable or alternating current is required. Give
the desired voltage, number of phases and frequency.
3. Say whether manual flow regulation can be used (with DC and very small AC plants) or if regulation
by an automatic governor is needed.

MEASURING GROSS HEAD


Method No. 1
1. Equipment <see figure 1>

a. Surveyor's leveling instrument--consists of a spirit level fastened parallel to a telescopic sight.


b. Scale--use wooden board approximately 12 ft in length.
2. Procedure <see figure 2>

a. Surveyor's level on a tripod is placed downstream from the power reservoir dam on which the
headwater level is marked.
b. After taking a reading, the level is turned 180[degrees] in a horizontal circle. The scale is placed
downstream from it at a suitable distance and a second reading is taken. This process is repeated until the
tailwater level is reached.

Method No. 2

This method is fully reliable, but is more tedious than Method No. 1 and need only be used when a
surveyor's level is not available.
1. Equipment <see figure 3>

a. Scale
b. Board and wooden plug
c. Ordinary carpenter's level
2. Procedure <see figure 4>

a. Place board horizontally at headwater level and place level on top of it for accurate leveling. At the
downstream end of the horizontal board, the distance to a wooden peg set into the ground is measured
with a scale.
b. The process is repeated step by step until the tailwater level is reached.

MEASURING FLOW
Flow measurements should take place at the season of lowest flow in order to guarantee full power at all
times. Investigate the stream's flow history to determine the level of flow at both maximum and
minimum. Often planners overlook the fact that the flow in one stream may be reduced below the
minimum level required. Other streams or sources of power would then offer a better solution.
Method No. 1

For streams with a capacity of less than one cubic foot per second, build a temporary dam in the stream,
or use a "swimming hole" created by a natural dam. Channel the water into a pipe and catch it in a bucket
of known capacity. Determine the stream flow by measuring the time it takes to fill the bucket.
Stream flow (cubic ft/sec) = Volume of bucket (cubic ft)
---------------------------Filling time (seconds)
Method No. 2
For streams with a capacity of more than 1 cu ft per second, the weir method can be used. The weir is
made from boards, logs, or scrap lumber. Cut a rectangular opening in the center. Seal the seams of the
boards and the sides built into the banks with clay or sod to prevent leakage. Saw the edges of the
opening on a slant to produce sharp edges ont he upstream side. A small pond is formed upstream from
the weir. When there is no leakage and all water is flowing through the weir opening, (1) place a board
across the stream and (2) place another narrow board at right angles to the first, as shown below.

Use a carpenter's level to be sure the second board is level.


Measure the depth of the water above the bottom edge of the weir with the help of a stick on which a
scale has been marked. <see figure 5> Determine the flow from Table 1 on page 56.

Table 1
FLOW VALUE (Cubic Feet/Second)
Weir Width
Overflow Height
3 feet
4 feet
1.0 inch
0.24
0.32
2.0 inches
0.67
0.89
4.0 inches
1.90
2.50
6.0 inches
3.50
4.70
8.0 inches
5.40
7.30
10.0 inches
7.60
10.00
12.0 inches
10.00
13.30

5 feet
0.40
1.06
3.20
5.90
9.00
12.70
16.70

6 feet
0.48
1.34
3.80
7.00
10.80
15.20
20.00

7 feet
0.56
1.56
4.50
8.20
12.40
17.70
23.30

8 feet
0.64
1.80
5.00
9.40
14.60
20.00
2 6.60

9 feet
0.72
2.00
5.70
10.50
16.20
22.80
30.00

Method No. 3
The float method is used for larger streams. <see figure 6> Although it is not as accurate as the previous
two methods, it is adequate for practical purposes.

Choose a point in the stream where the bed is smooth and the cross section is fairly uniform for a length
of at least 30 ft. Measure water velocity by throwing pieces of wood into the water and measuring the
time of travel between two fixed points, 30 ft or more apart. Erect posts on each bank at these points.
Connect the two upstream posts by a level wire rope (use a carpenter's level).
Follow the same procedure with the downstream posts. Divide the stream into equal sections along the
wires and measure the water depth for each section. In this way, the cross-sectional area of the stream is
determined. use the following formula to calculate the flow:
Stream Flow (cu ft/sec) = Average cross-sectional flow area (sq ft) X velocity (ft/sec)

MEASURING HEAD LOSSES


"Net Power" is a function of the "Net Head." The "Net Head" is the "Gross Head" less the "Head Losses."
The illustration below shows a typical small water power installation. The head losses are the openchannel losses plus the friction loss from flow through the penstock. <see figure 7>

Open Channel Bead Losses


The headrace and the tailrace in the illustration above are open channels for transporting water at low
velocities. The walls of channels made of timber, masonry, concrete, or rock, should be perpendicular.
Design them so that the water level height is one-half of the width. Earth walls should be built at a
45[degrees] angle. Design them so that the water level height is one-half of the channel width at the
bottom. At the water level the width is twice that of the bottom. <see figure 8>

The head loss in open channels is given in the nomograph. The friction effect of the material of
construction is called "N". Various values of "N" and the maximum water velocity, below which the
walls of a channel will not erode are given.

TABLE II
Maximum Allowable Water Velocity
Material of Channel Wall (feet/second) Value of "n"
Fine grained sand
0.6
0.030
Course sand
1.2
0.030
Small stones
2.4
0.030
Coarse stones
4.0
0.030
Rock (Smooth)
25.0
0.033
Rock (Jagged)
25.0
0.045
Concrete with sandy water
10.0
0.016
Concrete with clean water
20.0
0.016
Sandy loam, 40% clay
1.8
0.030
Loamy soil, 65% clay
3.0
0.030
Clay loam, 85% clay
4.8
0.030
Soil loam, 95% clay
6.2
0.030
100% clay
7.3
0.030
Wood
0.015
Earth bottom with rubble sides
0.033
The hydraulic radius is equal to a quarter of the channel width, except for earth-walled channels where it
is 0.31 times the width at the bottom.
To use the nomograph, a straight line is drawn from the value of "n" through the flow velocity to the
reference line. The point on the reference line is connected to the hydraulic radius and this line is
extended to the head-loss scale which also determines the required slope of the channel.

Using a Nomograph
After carefully determining the water power site capabilities in terms of water flow and head, the
nomograph is used to determine:
The width/depth of the channel needed to bring the water to the spot/location of the water turbine.
The amount of head lost in doing this.

To use the graph, draw a straight line from the value of "n" through the flow velocity through the
reference line tending to the hydraulic radius scale. The hydraulic radius is one-quarter (0.25) or (0.31)
the width of the channel that needs to be built.

In the case where "n" is 0.030, for example, and water flow is 1.5 cubic feet/second, the hydraulic radius
is 0.5 feet or 6 inches.
If you are building a timber, concrete, masonry, or rock channel, the total width of the channel would be 6
inches times 0.25, or 2 feet with a depth of at least 1 foot.
If the channel is made of earth, the bottom width of the channel would be 6 times 0.31, or 19.5 inches,
with a depth of at least 9.75 inches and top width of 39 inches.
Suppose, however, that water flow is 4 cubic feet/second. Using the graph, <see graph> the optimum
hydraulic radius would be approximately 2 feet--or for a wood channel, a width of 8 feet. Building a
wood channel of this dimension would be prohibitively expensive.

However, a smaller channel can be built by sacrificing some water head. For example, you could build a
channel with a hydraulic radius of 0.5 feet or 6 inches.
To determine the amount of head that will be lost, draw a straight line from the value of "n" through the
flow velocity of 4 [feet3]/second to the reference line.
Now draw a straight line from the hydraulic radius scale of 0.5 feet through the point on the reference line
extending this to the head-loss scale which will determine the slope of the channel. In this case about 10
feet of head will be lost per thousand feet of channel. If the channel is 100 feet long, the loss would only
be 1.0 feet--if 50 feet long, 0.5 feet, and so forth.

Pipe Bead Loss and Penstock Intake


The trashrack consists of a number of vertical bars welded to an angle iron at the top and a bar at the
bottom (see Figure below).

The vertical bars must be spaced so that the teeth of a rake can penetrate the rack for removing leaves,
grass, and trash which might clog up the intake. Such a trashrack can easily be manufactured in the field
or in a small welding shop. Downstream from the trashrack, a slot is provided in the concrete into which a
timber gate can be inserted for shutting off the flow of water to the turbine.

The penstock can be constructed from commercial pipe. The pipe must be large enough to keep the head
loss small. The required pipe size is determined from the nomograph. A straight line drawn through the
water velocity and flow rate scales gives the required pipe size and pipe head loss. Head loss is given for
a 100-foot pipe length.
For longer or shorter penstocks, the actual head loss is the head loss from the chart multiplied by the
actual length divided by 100. If commercial pipe is too expensive, it is possible to make pipe from native
material; for example, concrete and ceramic pipe, or hollowed logs. The choice of pipe material and the
method of making the pipe depend on the cost and availability of labour and the availability of material.

APPENDIX II
SMALL DAM CONSTRUCTION
This appendix is not designed to be exhaustive; it is meant to provide background and perspective for
thinking about and planning dam efforts. While dam construction projects can range from the simple to
the complex, it is always best to consult an expert, or even several; for example, engineers for their
construction savvy and an environmentalist or concerned agriculturalist for a view of the impact of
damming.

Introduction to:

Earth Dams
Crib Dams
Concrete and Masonry Dams

EARTH DAMS

An earth dam may be desirable where concrete is expensive and timber scarce. It must be provided with a
separate spillway of sufficient size to carry off excess water because water can never be allowed to flow
ovewr the crest of an earth dam. Still water is held satisfactorily by earth but moving water is not. The
earth will be worn away and the dam destroyed.
The spillway must be lined with boards or concrete to prevent seepage and erosion. The crest of the dam
may be just wide enough for a footpath or may be wide enough for a roadway, with a bridge placed
across the spillway.
The big problem in earth-dam construction is in places where the dam rests on solid rock. It is hard to
keep the water from seeping between the dam and the earth and finally undermining the dam.
One way of preventing seepage is to blast and clean out a series of ditches, or keys, in the rock, with each
ditch about a foot deep and two feet wide extending under the length of the dam. Each ditch should be
filled with three or four inches of wet clay compacted by stamping it. More layers of wet clay can then be
added and the compacting process repeated each time until the clay is several inches higher than bedrock.
The upstream half of the dam should be of clay or heavy clay soil, which compacts well and is
impervious to water. The downstream side should consist of lighter and more porous soil which drains
quickly and thus makes the dam more stable than if it were made entirely of clay.

CRIB DAMS

The crib dam is very economical where lumber is easily available: it requires only rough tree trunks, cut
planking, and stones. Four- to six-inch tree trunks are placed 2-3 feet apart and spiked to others placed
across them at right angles. Stones fill the spaces between timbers. The upstream side (face) of the dam,
and sometimes the downstream side, is covered with planks.
The face is sealed with clay to prevent leakage. Downstream planks are used as an apron to guide the
water which overflows the dam back into the stream bed. The dam itself serves as a spillway in this case.
The water coming over the apron falls rapidly. Prevent erosion by lining the bed below with stones. The
apron consists of a series of steps for slowing the water gradually.
Crib dams must be embedded well into the embankments and packed with impervious material such as
clay or heavy earth and stones in order to anchor them and to prevent leakage. At the heel, as well as at
the toe of crib dams, longitudinal rows of planks are driven into the stream bed. These are priming planks
which prevent water from seeping under the dam, and they also anchor it.
If the dam rests on rock, priming planks cannot and need not be driven; but where the dam does not rest
on rock they make it more stable and watertight. These priming planks should be driven as deep as
possible and then spiked to the timber of the crib dam.
The lower ends of the priming planks are pointed as shown in the Figure on page 69 and must be placed
one after the other as

shown. Thus each successive plank is forced, by the act of driving it closer against the preceding plank,
resulting in a solid wall. Any rough lumber may be used. Chestnut and oak are considered to be the best
material. The lumber must be free from sap, and its size should be approximately 2" X 6".
In order to drive the priming planks, considerable force may be required. A simple pile driver will serve
the purpose. The Figure below shows an excellent example of a pile driver.

CONCRETE AND MASONRY DAMS


Concrete and masonry dams more than 12 feet high should not be built without the advice of an engineer
with experience in this field. Dams require knowledge of the soil condition and bearing capacity as well
as of the structure itself.

A stone dam can also serve as a spillway. It can be up to 10 feet in height. It is made of rough stones. The
layers should be bound by concrete. The dam must be built down to a solid and permanent footing to
prevent leakage and shifting. The base of the dam should have the same dimensions as its height to give it
stability.

Small concrete dams should have a base with a thickness 50 percent greater than height. The apron is
designed to turn the flow slightly upwards to dissipate the energy of the water and protect the downstream
bed from erosion.

APPENDIX III
PUMP SELECTION
Design for a Simple Pump
One choice for a water-powered pump is a positive displacement pump. Such pumps are called by various
names: bucket pump, lift pump, piston pump, windmill pump, and occasionally even simply by brand
name, such as "Rocket" pump. Numerous models are available commercially and vary in cost from a few
dollars for small capacity pumps to several hundred for high capacity, high head, durable, well
manufactured units. However, pumps can be manufactured at low cost in the simplest of workshops.
One single acting pump attached to the wheel will cause speed surges on the wheel because actual
pumping takes place only half the time, while the other half is spent filling the cylinder. During the filling
stage, considerably less wheel torque is required than when pumping is being done. The speed surge can
be partially overcome by using:
* Two single-acting pumps 180[degrees] out of phase so that one of the pumps is always doing useful
work.

* A double-acting pump which has the same effect as the one above but is built in one unit; or

* best of all, two double-acting pumps 90[degrees] out of phase.

Use of multiple simple pumps improves the overall efficiency of the system. (In general, one unit can be
attached easily to a crank at each end of the wheel shaft.)
Table 1.
Quantities of Water Pumped Per Stroke for Single-Acting Pumps of Various Bore and Stroke Sizes
(Imperial Gallons)
Stroke (in.)
Bore (in.)
1-1/4
1-1/2
2
2-1/2
3
3-1/2
4
5

2-1/4

4
.009
.013
.023
.035
.052
.070
.092
.143

6
.016
.023
.040
I -064
.092
.125
.163
.255

8
.023
.035
.062
.095
.139
.187
.245
.382

10
.032
.045
.082
.127
.184
.248
.227
.510

12
.040
.057
.102
.159
.230
.312
.410
.638

.049
.069
.122
.191
.278
.276
.489
.765

DESIGN FOR A SIMPLE PUMP


An Easily Constructed Piston Pump
This pump <see figure> was designed by P. Brown (of the Mechanical Engineering Workshop at the
Papua New Guinea University of Technology) with a view to manufacture in Papua New Guinea.
Consequently the pump can be built using a minimum of workshop equipment--most parts are standard
pipe fittings available from any plumbing supplier.

A PVC pipe can be used in place of copper pipe. This eliminates the need for a pipe reducer. The PVC
pipe can have a uniform diameter throughout.
To avoid having to bore and hone a pump cylinder, a length of copper or PVC pipe is used. If care is
taken to select an undamaged length of pipe and to see that the pipe is not damaged during construction,
this system has proved quite satisfactory.
As can be seen from the cross-sectional diagram, the ends of the pump body consist of copper pipe
reducers silver-soldered onto the pump cylinder. This does make disassembly of the pump difficult, but
avoids the use of a lathe.
If a lathe is available, a screwed end could be silver-soldered to the upper end of the pump to allow for
simple disassembly.
The piston of the pump consists of a 1/2" thick PVC flange with holes drilled through it (see diagram on
page 78). A leather bucket is attached above the piston and together with the holes serves as a non-return
valve. In this type of pump the bucket must be made of fairly soft leather, a commercial leather bucket is
not suitable. Bright steel bar is used as the drive rod. Threads must be cut into the ends of the rod with a
die.
A galvanized nipple is silver-soldered to the top copper reducer of the pump to allow the discharge pipe
to be attached.
An `O' ring seal of the type used to join PVC pipe is used as a seal for the foot valve. This seal does not
require any fixing since it push fits into the lower copper pipe reducer. A 1/2" screwed flange with a plug
in its center forms the plate for the foot valve. This plate is prevented from rising up the bore of the pump
by three brass pegs fitted in through the sidewall of the pump above the valve plate. Silver-solder the pegs
to prevent leakage or movement.
Parts and tools for a 4" bore X 9" stroke pump include the following:
Parts
1 12" X 4" dia copper tube
2 4" to 1/2" copper tube reducers

1 1-1/2" galvanized nipple


1 1/2" screwed flange
1 1/2" plug
1 1/2" PVC flange
1 rubber `O' ring, 4" dia
1 4-1/2" dia piece of leather
1 15" X 1/2" dia bright steel bar
1 1/8" dia brazing rod

Tools

Handi gas kit


Silver solder
Hand drill
1/2" Whitworth die
1/2" Whitworth tap
Hacksaw
Hammer

APPENDIX IV
CALCULATING BEARING AND SHAFT SIZES
CALCULATING BEARING SIZE
Because it is very likely that people using this material will want to change the size of the waterwheel
they construct, the following information is provided to serve as a basis for determining the size of the
bearings which must be used.
Approximate Weight Carried by Each Bearing Excluding Loads Due to Attached Machinery (per
Meter of Width of the Wheel) (kg)
11
14.5
23
5
16
21.5
32
43
54.5
7.5
20
27.3
40.5
57
73
10
39
64
84
107
152
214
15
82
109
139
200
307
20
132
168
241
348
25
150
202
289
418
30
373
552
40
464
682
50
800
60
Bearing diameters required to support the various loads are given in the table on the following page
calculated on the basis of 100 psi (i.e., a hardwood such as oak) in parallel usage and 200 psi for end
grain usage. Values are given to 90.90 kgs to allow for the largest reasonable bearing loads.
(*) Outside wheel diameter minus inside wheel diameter divided by 2.
Minimum Bearing Inside Diameter Required For Various Loadings (cm)
Load (kg)
45.5 91 227 454 908
2272 4545 9090
2.5 3.8 5.75 8.25 10.88 17.75 25.5 35.5
Parallel Usage
1.5 2.5 4.5 5.75 8.25 12.5 17.75 25.5
End Grain Usage
These bearings are assumed to be steel on wood. It is likely that with metal shafts used in the larger sizes
of waterwheels, the bearing will be considerably larger than the required shaft size. A "built-up and
banded" bearing may be used. This is accomplished by attaching a wooden cylinder to the wheel shaft at
the bearing location to bring the cylinder's outside diameter to the necessary size. Then steel bands are
bent and fastened to the cylinder.

Calculating Shaft Size


Waterwheel shafts may be made of wood or steel. The diameter of the shaft depends on the material used
and the dimensions of the wheel. The tables below give minimum shaft diameters for bearing loads up to
45.45 kgs.
Minimum Standard Pipe Sizes for Use as Axles With Bearings at 30cm From Wheel Edge (Metal
Shafts)
45.5 91
227 454 908 2270 4540
Bearing Load (kg)
2.5
3.75 6.25 7.5 10 15
20
Pipe Diameter cm) Solid Metal Shaft
Minimum Standard Hardwood Sizes for Use as Axles With Bearings at 30cm From Wheel Edge
(Wooden Shafts)
45.5
91
227 454 908 2270
4540
Bearing Load (kg)
3.75
6.25
9
18
33
86.5
173
Wood Shaft Diameter (cm)
When comparing these figures with the bearing diameters, it can be seen that for pipe or a solid steel
shaft, a wooden bearing will need to be built up. With wooden shafts, the required shaft diameter will
usually exceed the required bearing diameter giving one the choice of reducing the shaft diameter at the
bearing location (but only there) or of using larger bearings.
In either case, the shaft must be banded with steel, sleeved with a piece of pipe, or given some similar
protection against wear in the bearing.

APPENDIX V
DECISION MAKING WORKSHEET
If you are using this as a guideline for using the Waterwheel in a development effort, collect as much
information as possible and if you need assistance with the project, write VITA. A report on your
experiences and the uses of this manual will help VITA both improve the book and aid other similar
efforts.
VOLUNTEERS IN TECHNICAL ASSISTANCE 1600 Wilson Boulevard, Suite 500 Arlington, Virginia
22209, USA

CURRENT USE AND AVAILABILITY

Describe current agricultural and domestic practices which rely on water at some point.
What water power sources are available? Include rivers, streams, lakes, ponds. Note whether
sources are small but fast-flowing, large but slow-flowing, etc.
What is water used for traditionally?
Is water currently being used to provide power for any purpose? If so, what and with what
positive or negative results?

Are there dams already built in the area? If so, what have been the effects of the damming? Note
particularly any evidence having to do with the amount of sediment carried by the water--too
much sediment can create a swamp.
If water resources are not now harnessed, what seem to be the limiting factors? Does the cost of
the effort seem prohibitive? Does the lack of knowledge of water potential limit its use?

NEEDS AND RESOURCES

Based on current agricultural and domestic practices, what seem to be the areas of greatest need?
Is power needed to run currently hand-powered machines such as grinders, saws, pumps?
What are the characteristics of the problems? Is the local population aware of the problem/need?
How do you know?
Has any local person, particularly someone in a position of authority, expressed the need or
expressed any interest in this technology/ If so, can someone be found to help the technology
introduction process?
Are there local officials who could be involved and tapped as resources?
How can you help the community decide which technology is appropriate for them?
Given water power sources available which water resources seem to be available and most useful?
For example, one stream which runs quickly year around and is located near to the center of
agricultural activity may be the only feasible source to tap for power.
Define water power sites in terms of their inherent potential for power generation. In other words,
one water source may be a power resource only if harnessed by an expensive turbine.
Are any materials for constructing water power technologies available locally? Are local skills
sufficient? Some water power applications demand a rather high degree of construction skill. Is
surveying equipment available? Do you need to train people?
Can you meet the following needs?
Some aspects of the waterwheel project require someone with experience in woodworking and surveying.
Estimated labour time for full-time workers is:
4 hours skilled labour
40 hours unskilled labour.
If this is a part-time project, adjust the times accordingly.
Do a cost estimate of the labour, parts, and materials needed.
Does the technology require outside funding? Are local funding sources available?
What is your schedule? Are you aware of holidays and planting or harvesting seasons which may
affect timing?
How will you spread information on, and promote use of, the technology?

IDENTIFY APPROPRIATE TECHNOLOGY


* Is more than one water power technology applicable? Weigh the costs of various technologies relative
to each other--fully in terms of labour, skill required, materials, installation and operation costs.
Remember to look at all the costs.
* Are there choices to be made between say a waterwheel and a windmill to provide power for grinding
grain? Again weigh all the costs: feasibility, economics of tools and labour, operation and maintenance,
social and cultural dilemmas.
* Are there local skilled resources to guide technology introduction in the water power area?
* Where the need is sufficiently large-scale and resources are available, consider a manufactured turbine
and a group effort to build the dam and otherwise install the turbine.
* Could a technology such as the hydraulic ram be usefully manufactured and distributed locally? Is there
a possibility of providing a basis for a small business enterprise?

FINAL DECISION
* How was the final decision reached to go ahead--or not go ahead--with this technology?

APPENDIX VI
RECORD KEEPING WORKSHEET
CONSTRUCTION
Photographs of the construction process, as well as the finished result, are helpful. They add interest and
detail that might be overlooked in the narrative.

A report on the construction process should include much very specific information. This kind of detail
can often be monitored most easily in charts (such as the one below). <see report 1>

Some other things to record include:


Specification of materials used in construction.
Adaptations or changes made in design to fit local conditions.
Equipment costs.
Time spent in construction--include volunteer time as well as paid labour; full- or part-time.
Problems--labour shortage, work stoppage, training difficulties, materials shortage, terrain,
transport.
OPERATION
Keep log of operations for at least the first six weeks, then periodically for several days every few
months. This log will vary with the technology, but should include full requirements, outputs, duration of
operation, training of operators, etc. Include special problems that may come up--a damper that won't
close, gear that won't catch, procedures that don't seem to make sense to workers, etc.
MAINTENANCE
Maintenance records enable keeping track of where breakdowns occur most frequently and may suggest
areas for improvement or strengthening weakness in the design. Furthermore, these records will give a
good idea of how well the project is working out by accurately recording how much of the time it is
working and how often it breaks down. Routine maintenance records should be kept for a minimum of six
months to one year after the project goes into operation. <see report 2>

SPECIAL COSTS
This category includes damage caused by weather, natural disasters, vandalism, etc. Pattern the records
after the routine maintenance records. Describe for each separate incident:
Cause and extent of damage.
Labor costs of repair (like maintenance account).
Material costs of repair (like maintenance account).
Measures taken to prevent recurrence.
.

MOTORS AS GENERATORS
FOR MICRO-HYDRO POWER
CONTENTS
PREFACE
ACKNOWLEDGEMENTS
INTRODUCTION
1. Advantages and disadvantages of induction generators
2. Induction machine construction and operation
Construction
Operation
Induction motor operation
Supply-connected induction generator operation
Stand-alone induction generator operation
3. Selection of an induction motor for use as a generator
Rotor type
Site conditions
Lubrication
Frame type
Insulation and temperature rise
Efficiency
Power rating
Voltage rating
Speed rating / drive arrangement
4. Excitation capacitor requirements
Voltage and current relationships in three-phase circuits
Capacitor connection
Selection of capacitors
Calculation of excitation capacitance
5. Operating voltage and frequency
Operating voltage
Operating frequency
Using the operating voltage and frequency to advantage
6. The effect of load upon generator output
Resistive load
Inductive load
Capacitive load
7. Single-phase output from a three-phase machine
8. Generator protection
Overload
Underload
Overspeed
9. Fixed and variable load systems
Fixed load systems
Variable load systems
Controller options
10. Motor starting
Techniques for improving motor starting

11. Generator commissioning


Safety
Output frequency and power
Phase rotation
Underload testing
Adjustment for maximum efficiency
Appendix 1: Three-phase circuit theory
Line and phase quantities
Voltage and current relationships
Power and power factor
Appendix 2: Induction generator efficiency
Appendix 3: Single transformer systems
Appendix 4: Halving the operating voltage of an induction machine
Appendix 5: Overvoltage trip
Appendix 6: Power factor correction for motor starting
PREFACE
Micro-hydro is a valuable source of energy for rural industries and village electrification schemes. It has
been a traditional method of grain processing throughout the world and played a major role in
modernization and industrial development in Europe and North America. Micro-hydro now offers
similar potential to most developing countries, with applications in village lighting, mechanized food
processing, and the supply of power to small-scale industrial activities.
This book contributes an important element of knowledge which is needed to realize this great
potential. In our efforts over the years to assist local manufacturers and installers of hydro equipment,
the staff of ITDG have found that conventional generators are difficult to obtain or are unreliable in
service. Motors, on the other hand, are always obtainable locally and are very robust in operation. This
book is intended to help local manufacturers and rural development engineers to select a motor and
convert it for use as a generator for a micro-hydro scheme.
The development of this technology is the fruit of collaborative efforts between ITDG staff and
colleagues and friends in Sri Lanka, Nepal and Indonesia, where the field work has been carried out. It
has also been an excellent example of what can be achieved by matching the resources of UK research
institutions to needs in developing countries.
The use of motors as generators is now well proven and promises to be an important element in
establishing self-sustaining local capacity for village-scale hydro in developing countries.
Take-up of the technology outlined in this book is now freely available to individuals, communities and
organizations overseas, helping to fulfil ITDG's strategic aim of wide dissemination of appropriate
technologies, which can be manufactured locally at affordable cost. ITDG continues to widen the
availability of appropriate technologies through its on-going programme of training courses in microhydro and other aspects of rural energy. Specific courses are also held on induction generators and local
manufacture of electronic controllers. For details, please write to ITDG at the Rugby address.
ITDG Energy Unit
Intermediate Technology Development Group (ITDG)
Myson House, Railway Terrace, Rugby CV21 3HT, UK
ACKNOWLEDGEMENTS
I would like to express my thanks to the Intermediate Technology Development Group (ITDG) and the
Overseas Development Administration (ODA) Engineering Division for funding publication of this book,
as well as much of the research and development work that has made it possible. In particular, their
open technology approach has made it possible for technical data to be documented and freely
transferred throughout the course of this work, and has included several training courses for local
engineers and manufacturers.
I would also like to acknowledge The Nottingham Trent University for providing research facilities and
technical and financial support.

In addition, I am grateful to Ian Macwhinnie, Arthur Williams, Keith Pratt and Doreen Smith for reading
the manuscript and making valuable suggestions.
Thanks are also due to Brook Crompton for permission to publish their motor data, the photograph of
an induction motor (Figure 1) and the technical drawing (Figure 4).
Nigel Smith Nottingham 1994
INTRODUCTION
Until recently, micro hydro installations have always used synchronous generators to produce a.c.
electricity if no grid supply is available. Often, for reasons of availability and low cost, machines designed
for use with petrol or diesel sets are used. These machines are generally designed for intermittent use,
under direct drive conditions. It is little wonder that they do not last long in continuous operation,
especially when belt driven by a water turbine with a runaway speed considerably higher than its
normal operating speed. Quality synchronous machines are built for such arduous duties, though they
are less easy to obtain and are expensive.
In recent years, induction motors have been installed as generators on micro hydro schemes. The
induction motor is the most common type of electrically powered prime-mover used by industry.
Induction generators are sometimes referred to as IGs or IMAGs - induction motors as generators. In
Nepal alone more than one hundred induction generators have been installed in the past six years, and
they have proven to be considerably more reliable than the available synchronous generators.
With early schemes, because of the absence of a suitable control system, a near constant load had to be
maintained at all times. This was generally achieved by having a single switch or contactor next to the
generator and permanently connecting all the loads to the supply. A low cost induction generator
controller (IGC) has now been developed, and locally manufactured IGCs are fitted to all but the
smallest of schemes. These allow a wide variety of loads to be connected and disconnected as required.
It has been assumed that the reader has a basic understanding of single and three-phase circuits and
induction motors. Those without the necessary background knowledge should read the relevant
chapters of standard diploma/certificate level electrical engineering books.
For consistency, a supply voltage of 240 V single-phase and 415 V three-phase and a frequency of 50 Hz
has been assumed throughout the book. However, the principles taught can be applied to all other
supply specifications.

1. ADVANTAGES AND DISADVANTAGES OF INDUCTION GENERATORS


The main advantages and disadvantages of using induction machines instead of synchronous generators
for stand-alone micro hydro are as follows.
Advantages
Availability
Induction motors are much more widely available than synchronous generators. In some cases, secondhand machines can be reconditioned in order to reduce costs.

Cost
Induction generators, including their excitation capacitors, are generally cheaper than synchronous
generators. This is especially true for low power ratings. For example, a 10 kW induction generator is
typically half the cost of a synchronous generator. This price difference varies between countries, since
it depends on whether local manufacture takes place and on the sources of imported machines. In some
countries there is little or no financial advantage to using induction generators for schemes of more
than 25 kW, whereas in others there is an appreciable cost saving.
Robustness
Induction machines are very robust and have a simple construction, as shown in Figure 1. They have no
winding, diodes or slip rings on their rotor. Solid, normally cast bars, replace the rotor winding and
enable the rotor to withstand considerable overspeed. In addition, the machines are normally totally
enclosed, ensuring good protection against dirt and water. They are designed for continuous operation
with belt drives under arduous industrial conditions.
Disadvantages
Voltage rating
Induction machines are not always available with suitable voltage ratings for use as generators.
Modification to the winding connections, or in extreme cases rewinding, may be required.
Calculation required
Whilst synchronous generators can be purchased ready for use, the induction machine will not work
without capacitors of a suitable value being fitted.
Motor starts
Motors are more easily started with synchronous generators than induction generators. Induction
motors, with a capacity that is large when compared to the generator rating, can cause severe voltage
dips or even loss of excitation when started from induction generators.
This book explains how to choose and if necessary modify the winding arrangement of an induction
machine in order to make it generate at the desired voltage, and how to calculate the capacitance
required. It also advises on ways of improving motor starting and the limits of motor starting capability.
It explains how to use a three-phase motor as a single-phase generator, and describes the protection
and control requirements.
2. INDUCTION MACHINE CONSTRUCTION AND OPERATION
Construction
The mechanical features of an induction machine can be seen from Figure 2. From an electrical point of
view, the induction machine consists of two parts: a fixed, wound stator core on the outside and a rotor
that rotates in the centre. The stator winding consists of coils of insulated copper wire fixed into slots in
the core to form a distributed winding of a similar type to that used with a synchronous generator.
The induction machine rotor is very different from that of a synchronous generator. The standard
squirrel cage rotor core is cylindrical and built up from thin sheets of steel into which slots have been
punched for the conductors. The conductors generally consist of aluminium bars that are short-circuited
at each end by aluminium rings. For small machines, the bars and end-rings are cast in one operation
using the rotor core as part of the die. For large machines, copper or aluminium bars are welded or
brazed to the end-rings.
Operation
In order to understand the operation of a stand-alone, i.e. non grid connected, induction generator it is
easiest to take motor operation as a starting point.
Induction motor operation
When an induction machine is connected to an a.c. supply, magnetizing current flows from the supply
and creates a rotating magnetic field in the machine.
The rotating field cuts the short-circuited rotor bars, inducing currents in them which, because they are
flowing in the magnetic field, react with it producing a torque. This torque drags the rotor round with
the field, but at a slightly lower speed. The small difference in speed arises because without it no

currents would be induced in the rotor and, therefore, no torque would be produced to turn it. When a
load is applied to the motor the speed difference will increase as a greater torque must be produced.

Ref.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

Part Description
Endshield, drive end, foot mounted
Grease nipple (when fitted)
Grease relief screw (when fitted)
Pre loaded washer
Oil seals
Bearing, drive end
Rotor assembly
Stator frame with or without feet
Eyebolt (when fitted)
Inside bearing cap, non-drive end
Bearing, non-drive end
Circlip
Endshield, non-drive end
Grease nipple (when fitted)
Inside bearing cap screws
Fan
Fan locating pin
Fan cover
Terminal box lid
Terminal box lid gasket
Terminal board
Terminal box
Terminal box gasket
Detachable feet
Drain plug (when fitted)
Face flange

27 'D' flange
28 Feet fixing screws and Belleville washer
The difference between the speed of the rotor and the speed of the rotating field is called the 'slip' and
is defined as:
(nc - n )
slip, s = ijd
(1)
ns
Where ns is synchronous speed (the speed of the rotating field)
nr is the rotor speed
Without load connected, the slip of the induction motor will be very small, less than 0.01 (that is 1%).
For a machine of 1 kW the full load slip will be about 0.05 (or 5%). Larger machines have smaller slips.
For more information see the speed column in Table 1.
Supply-connected induction generator operation
If the same supply-connected induction machine is now driven at above synchronous speed, so that the
slip becomes negative (n, > ns), a torque is supplied to the rotor rather than by the rotor and the
machine acts as a generator, supplying power to the network. However, it still takes its magnetizing
current from the supply in order to create the rotating field, just as though it were a motor. The full load
power output is achieved at a slip of similar value (but negative) to the full load motoring slip. Example 1
illustrates the difference in slip and shaft speed between motor and generator operation.
Stand-alone induction generator operation
The magnetizing current of an induction machine can be supplied, in total or in part, by capacitors. In
fact, capacitors are often fitted to large induction motors and supply-connected induction generators to
reduce the reactive current drawn from the supply, especially when the electricity company imposes
charges for poor power factor.
In the case of a stand-alone induction generator, the capacitors are the only external source of
magnetizing current. Therefore, in order to obtain the required operating voltage at the desired
frequency, the amount of capacitance must be carefully chosen.
Power Factor
Full load speed
FLC* (A) at 415 Efficiency
kW Poles
(rpm)
V
FL
3/4L
1/2L
FL
3/4L
1/2L
74 73 76 72 76 69 .79 .70
.51 .47
0.55 24 6 28101400900
1.32 1.50 1.65
.68 .61 54
73
72
69
.64
.42
78 77 78 77 74 74 .77 .70 .73 .60 .61 .48
0.75 24 6 28501400920
1.75 1.95 2.40
71
70
65
.62
.53
.41
79 79 78 78 73 71 .80 .65
1.1 2 4 6 2850 1410 900
2.5 3.0 3.1
.74 55 .58 64 43 .47
75
75
74
.66
76 82 78 83 70 80 .86 .71
1.5 2 4 6 28501420940
3.2 3.6 4.3
.82 .61 55 .73 .47 44
75
74
72
.65
81 80 81 81 80 80
.73 .58
22 2 4 6 28501420945
4.4 4.9 5.9
.86 .78 68 82 .71 58
77
77
75
.47
83 81 83 82 81 80
.70 .61
3.0 2 4 6 28601420950
5 8 6.5 6.4
.88 80 .78 83 74 .73
83
84
82
.61
85 83 86 84 85 83 .90 .80
40 2 4 6 28401420960
7.2 8 4 9.2
88 .72 .62 81 .61 .50
84
83
82
.72
85 85 85 86 83 85
.83 .74 .76 .61
5.5 2 4 6 2900 1445 955
106 11.3 12.8
.85 .80.71
84
85
83
.63
.50
86 86 86 86 85 85 .87 .83
7.5 2 4 6 29001450970
14.1 14.6 156
.84 76 .70 .77 64 .55
87
87
85
.77
87 88 87 88 85 87 .87 .86 .83 .81
11 2 4 6 2930 1460 970
20 20 24
74 .70 55
88
88
87
.73
.65

88 90 87 89 .89 .87
.85 .68
88 .80 .80
88
87
.83
.70
90 89 87 88 .89 .89 .87 .88 .75 .83
18.5 2 4 6 2950 1460 970
32 33 36
90
88
.81
.76
.67
88 91 86 90
22 2 4 6 2920 1460 970
38 38 41
.92 89 83 91 86 7B 88 81 .70
91
89
88 91 86 89 .89 .86
30 2 4 6 2945 1470 970
53 54 55
8991 91
87 83 .78 79 76 .70
91
90
.83
90 91 89 91 86 89
.85 .80 .80 .72
37 2 4 6 29501465975
64 66 68
89 .86 .83
92
91
89
.78
.70
90 91 89 91 87 89
.90 .81 .84 .74
45 2 4 6 29501465975
77 79 81
.91 .87 84
92
92
90
.79
.71
90 91 89 91 87 89
.90 .78 .84 .68
55 2 4 6 29551470985
93 98 96
91 .83 .86
92
92
90
.81
.74
Table 1 Motor performance data (Brook Crompton)
* Where FLC is the full load current.
Example 1
For the 4-pole 7.5 kW induction machine listed in Table 1:
a) What is the full load slip?
b) At what shaft speed will the machine operate as a generator when running at full load and 50 Hz?
a) The full load motor shaft speed is given as 1450 rpm. The synchronous speed is calculated from:
120 xf
15 24 6 2940 1470 970

27 27 29

88 90
88
90 89
90
88 90
91

(2)

b) Since full load generating slip is approximately equal in magmtude to full load motoring slip, but
negative, s = -0.033
Re-arranging Equation 1, nr = ns(l- s)
nr = 1500 (l-[-0.033]) = 1550rpm
Hence, instead of the shaft speed being 50 rpm below the synchronous speed, as a generator the shaft
speed is 50 rpm above the synchronous speed.

For a build-up of voltage to occur, sufficient remanent magnetism must be present in the rotor.
Remanent magnetism is the initial magnetism present in the rotor steel. It is generally sufficient to
produce a small voltage, of about a volt, at synchronous speed with no capacitance connected. There
may be insufficient magnetism if, since it was last used, the machine received a large impact or
generation was collapsed with a resistive load connected. Remanent magnetism is dependent upon the
type of steel used. Low loss steels, as used in energy efficient induction machines, tend to have the
lowest levels of remanent magnetism.
If there is insufficient magnetism to excite the generator at rated frequency, increase the generator
speed, because at higher frequency less magnetism is required for excitation to occur. In almost all
cases, this will be sufficient to excite the generator. However, if this fails, remanent magnetism can be
increased by connecting a d.c. supply for a few seconds across any two of the machine terminals, before
running the machine up to speed. A car battery is more than sufficient for this purpose. Ordinary dry

cells, connected in series, can also be used. High capacity dry cells, such as those used in large torches,
are best as they can source the high current required.
A simplistic but useful way to understand the basic operation of a stand-alone induction generator is to
represent the machine simply by its magnetizing reactance. The highly simplified equivalent circuit
shown in Figure 3 can then be used. This is a fairly accurate representation for the purpose of
determining capacitor requirements.
Figure 3 Highly simplified induction generator equivalent circuit

With the shaft rotating, current will begin to flow due to the remanent magnetism present in the rotor.
The capacitor current, Ic, will equal the magnetizing current, Im, and the machine and capacitors will act
as a resonant circuit at an angular frequency, a>, fixed by the shaft speed of the machine. Provided that
sufficient capacitance is present, the current will rapidly increase until stable operation is reached when
the impedance of the capacitors equals the magnetizing impedance as given by Equation 3.

Phase Current
Figure 4 No load excitation characteristic for an induction generator driven at constant speed (for three
values of excitation capacitance)
Stable operation occurs because the magnetizing inductance is a non-linear function of current, due to
magnetic saturation of the rotor and stator steel. Provided that sufficient capacitance is connected, the
voltage against current characteristic for the capacitor(s) meets the voltage against current
characteristic for the magnetizing inductance, as
shown in Figure 4. This is the operating point of the generator. Increasing the capacitance will increase
the operating voltage but, since more current flows into the machine, additional power will be lost as
heat in the stator windings.
3. SELECTION OF AN INDUCTION MOTOR FOR USE AS A GENERATOR
The number of options available when selecting an induction machine will depend upon the accessibility
of manufacturers and suppliers and the range of machines that they stock. The following guidelines
cover the broadest range of options that may be open to the purchaser, though some of the options are
rarely available.
Rotor type

Induction motors with cage rotors are by far the most common type of machine and should always be
selected. Some manufacturers also produce wound rotor machines, but these are more expensive and
less robust and should, therefore, be avoided. Wound rotor machines can be recognized by the fact that
they have additional terminals which connect to the rotor by means of slip rings.
Site conditions
Motors should have an IP classification on their nameplate which indicates their level of protection
against penetration by solids and liquids. The first number refers to penetration by solids and the
second to penetration by liquids. The levels of protection, or IP numbers, most commonly available are:
IP 21 Protected against solids greater than 15 mm and against
vertically falling dripping water. IP 22 Protected against solids greater than 15 mm and against dripping
water falling at any angle up to 15 from the vertical. IP 23 Protected against solids greater than 15 mm
and against water falling as a spray at any angle up to 60 from the vertical. IP 44 Protected against
solids greater than 1 mm and water splashing from any direction. IP 54 Protected against dust and
against water splashing from any direction. IP 55 Protected against dust and against water jets from any
direction.
Most motors are classified as Totally Enclosed Fan Ventilated (TEFV) and are protected to IP54 or IP55.
These machines are suitable for use in dusty environments, such as mills. For direct-drive applications
IP55 motors are best because of the additional protection that they provide.
Induction motors with a capacity above 10 kW are sometimes available with drip-proof frames (EP21,
IP22 or IP23). They are cheaper than TEFV machines and are, therefore, worth considering, except in
dusty environments.
Lubrication

0
10
20
30
Motor kW rating
Figure 5 Typical relubrication intervals for good quality machines
As shown in Figure 5, induction machine bearings do not require frequent lubrication. The relubrication
interval depends upon the speed and power rating of the machine. Because relubrication intervals are
long, manufacturers often do not fit regreasing facilities to their motors. This is acceptable if the
machine is run for only a few hours per week but, for generators operating continuously, the machine
will need to be
dismantled to regrease the bearings at regular intervals. It is best to avoid dismantling the machine,
especially in dusty environments, and much easier to apply grease using a grease gun. For these
reasons, motors fitted with regreasing facilities should be used whenever possible. Grease relief should
also be specified to prevent overgreasing.

Manufacturer's instructions for regreasing should be followed if these are available. Otherwise
relubrication should be carried out at the intervals shown in Figure 5 using the following procedure:
1) Wipe clean the grease gun fitting and regions around the motor grease fittings.
2) Remove the relief plugs.
3) To each bearing add a small quantity of grease compatible with the bearing size.
4) Allow the motor to run for about ten minutes before refitting the relief plugs so that excess grease
can be expelled.
Frame type
Manufacturers often provide the option of either foot or flange mounting, as shown in Figure 2. For
certain applications such as the direct-drive pelton, shown in Figure 6, flange mounting may be
preferable.
Insulation and temperature rise
The lifetime of the machine windings is dependent upon the number of hours per day that the machine
is used, the quality of the insulation and the operating temperature. A rough guide to the effect of
temperature upon insulation life is that for every 10C reduction in operating temperature the
insulation life of the windings will be doubled. The most common classifications of insulation in current
use are:
Class B permitting a winding temperature rise of 80C above a 40C ambient, or 120C maximum. Class F
permitting a winding temperature rise of 100C above a 40C ambient, or 140C maximum. Class H
permitting a winding temperature rise of 110C above a 40C ambient, or 150C maximum.
If the windings are operated at the maximum temperature for their classification, then, according to
insulation standards, the life of the insulation will be a minimum of 20,000 hours. In practice the life is
much longer, since the machine is unlikely to operate at its maximum temperature rise and in maximum
ambient temperature all the time.
Improvements in insulation over the years have meant that machines can run hotter, allowing thinner
conductors to be used. As a result, machines have become smaller and cheaper. However, using thinner
conductors reduces efficiency and nowadays it is the trade-off between efficiency and machine cost,
rather than maximum operating temperature, that tends to determine conductor size.
The cost of class F insulated wire is now little more than for class B. Hence, many modern induction
machines are wound with class F wire even if the operating temperature does not necessitate its use.
Indeed, some manufacturers use class F insulation but state that their motors are designed to operate
within class B limits. Such motors will have approximately four times the winding life of the same
machines fitted with class B insulation.
Since most induction generators are used continuously and down time for rewinding is very
inconvenient, insulation life is an important consideration and motors with high temperature insulation
should be sought.
In rare cases the generator output will need to be derated because of operation at high altitude.
Standard ratings apply to temperatures not exceeding 40C at altitudes up to 1000 metres above sea
level. The following correction factors should be used at higher altitudes.
Altitude
2000m 3000m 4000m
When ambient temperature is 40C reduce rated output by: 92% 85% 75%
Maximum ambient temperature for full rated output
32C 24C 16C
Table 2 Output derating for high altitude operation
Efficiency
Induction machines have a slightly lower efficiency when operated as generators than as motors, as
shown in Appendix 2. Often the only option for improving efficiency is to change the operating voltage
and frequency, as discussed in Chapter 5. However, some manufacturers produce a range of high
efficiency motors by increasing the amount of copper used in the windings and the quantity and/or
quality of the steel used in the stator and rotor cores.

Where available, these machines should be considered when selecting a suitable generator. They are
typically 50% more expensive than a standard motor, but this cost is often justified by the increased
power output, as shown in Example 2. Energy efficient machines have better efficiencies than standard
machines at part-load as well as full load.
The low-loss steels used in energy efficient machines result in lower remanent magnetism. To provide
sufficient magnetism for excitation to occur a battery may well be required, as explained in the previous
chapter.
Power rating
The greater the power drawn from the generator, the higher its operating temperature and hence the
shorter the life of its windings. To ensure long winding life when used as a generator, the machine
should be kept below its full load operating temperature as a motor.
Winding temperatures for an induction machine operating as a motor and as a generator are given in
Appendix 2. By careful choice of the operating conditions in generator mode, namely voltage and
frequency, a generator mode power rating similar to the motor mode power rating can be achieved
with an equivalent temperature rise. However, this assumes balanced operation and near optimum
operating conditions.
Since maintaining an equal load on each phase is rarely possible on small generating systems and
operation under optimum conditions is not always possible, induction machines should be derated
when used as generators. A derating factor of 0.8 is recommended and is also applicable to single-phase
generation from a three-phase machine, as explained in Chapter 7.
A de-rating factor of 0.8 provides a good safety margin, since heat transfer between the stator windings
helps to correct for temperature differences due to unequal currents in the windings.
Example 2

Avoid using a generator larger than required, as induction machines, especially of smaller capacities,
have poor part load efficiencies, as shown in Appendix 2.
Voltage rating
When selecting an induction machine the voltage rating is a very important consideration. In some cases
it will be possible to obtain a machine that can be used directly as a generator. In other cases some
modification to the windings will be necessary.
Small generators
Induction machine manufacturers often wind machines of up to 3 kW capacity to provide the choice of
either 240 V delta or 415 V star operation. These machines will be labelled 240/415 V on the nameplate.
The star and delta winding configurations are shown below. The 240 V connection will be used when
operating the machine as a single-phase generator, as explained in Chapter 7. The 415 V connection is
suitable for three-phase generation, as the presence of a star point, which can be earthed and forms the
neutral connection, enables single-phase (240 V) loads to be connected between any line and neutral,
and three-phase loads to be connected to the three lines.
U1

Motors that can be connected for either 240 V or 415 V operation are supplied with all six ends of their
windings available in the terminal box. This enables easy connection in star or delta as shown in Figure
8. Some small motors are supplied with the windings in star and with the star connection made
internally so that just three connections are provided at the terminal box.
In such cases, it is a simple job for a motor repairer to provide a connection to the star point at the
terminal box. If a spare terminal is not available a good earthing point to the machine case, within the
terminal box, can be used for the star point and shared with the earth connection. Reconfiguration in
delta again requires internal reconnection of the windings and the delta connected lines must be
brought to three terminals in the terminal box. Motors wound for 415 V delta connection should be
avoided, as explained in the next section.

Figure 8 Star and delta connections at the terminal block


Larger generators
Induction motors with a capacity above 3 kW are generally wound for star-delta starting. The machines
are designed to run with a 415 V delta connection as explained in Chapter 10.
From a generating point of view, a 415 V delta connection has the disadvantage that there is no star
point to serve as a neutral and earth connection. Hence, single-phase (240 V) loads cannot be connected
and there are significant disadvantages in terms of earthing. This winding arrangement can be used to
advantage in situations where there is a long transmission line, as explained in Appendix 3. However, for
most schemes 415 V delta is unsuitable and a star connected machine should be used.
There are three possibilities for obtaining a 415 V star connected machine. These, in order of ease and
cost, are as follows:
1) Deal directly with the manufacturer or his agent in order to purchase a machine wound for 240/415 V
operation. Most manufacturers will supply such machines at little or no extra cost, though the delivery
time is usually increased.
2) Reduce the operating voltage by reconnecting the groups of winding coils to a parallel configuration.
This can be achieved because machines are generally wound with an even number of coil groups per
phase, connected in series. Reconnection in parallel to yield 207.5 V, as explained in Appendix 4, is a
relatively straightforward operation, since only the ends of the coil sets need reconnecting and no
rewinding is required. To operate the machine at 240 V, an increase in operating frequency of

approximately 10% will be required to compensate for the increased saturation, as explained in Chapter
5.
3) Have a 415 V delta connected machine rewound for 415 V star. This is expensive as it costs
approximately half the price of a new machine. It is a good option if reconditioning a second-hand
machine that must be rewound anyway. Make sure that you use a rewinder who can scale the turns and
cross-sections of the cables correctly as many are used to just copying the existing winding.
If the first or third options are taken then it is worth considering winding the machine for a higher
voltage in order to improve efficiency, as explained in Chapter 5.
Speed rating / drive arrangement
Ideally the generator should be directly driven by the turbine. This has advantages in terms of increased
efficiency, reduced drive system costs, lower maintenance and simpler installation. Unfortunately,
turbine speeds are often much slower than standard generator speeds and therefore this approach is
not always possible. It may be suitable for high head or low flow installations or where high specific
speed turbines such as pumps as turbines are appropriate.

Induction machines have the advantage over synchronous generators that 6 pole and sometimes 8 pole
machines are available. However, for a given power rating, the cost relative to that of a 4 pole machine
is typically 1.5 times for a 6 pole machine and 2 times for an 8 pole machine. There is generally little
price difference between 2 pole and 4 pole machines.
With most crossflow installations a 4 pole machine is belt driven by the turbine, as shown in Figure 9.
For very slow speed turbines, where a single stage belt drive is insufficient, geared induction machines,
where available, are worth considering. Heavy duty gear boxes should be selected in order to ensure
good reliability. Figure 10 shows a 20:1 geared 4 pole induction motor which is used with a water wheel
in England. A tractor differential, already on site, is used to provide the first stage of the drive, though it
would have been possible to purchase a geared motor for the full 120:1 step up required.
Although standard induction machines are designed for belt drive, care must be taken to limit the radial
load in order to ensure good bearing life. Where possible, the motor supplier should be contacted for
information on pulley dimension limits and the belt manufacturer's catalogue consulted for belt tension
requirements.
4. EXCITATION CAPACITOR REQUIREMENTS
Voltage and current relationships in three-phase circuits
A good understanding of the voltage and current relationships in star and delta connected circuits is
required in order to calculate excitation capacitance values correctly. The essential terms and
relationships are given in Appendix 1.
Capacitor connection
For a three-phase generating system, the capacitors can be either star or delta connected, as shown in
Figure 11. In the case of the star connected system, the star point of the capacitors should not be
connected to the generator and system neutral as waveform distortion and increased losses will occur.

Figure 11 Star and delta connection of excitation capacitance


The star connection and delta connection are related as follows:
VCi =V3.VCs and ICa =\/S

1
C,
Since C =-------, CA =^(5)
co.Xc ______3_
If the capacitors are connected in star, then three times as much capacitance is required than for delta
connection, though lower voltage and therefore cheaper capacitors may be used. A sample calculation
is presented in Example A2 of Appendix 1.
Selection of capacitors
Generally the most suitable type of capacitor to use is the 'motor run' type, used with some singlephase induction motors. They are widely available in sizes up to and sometimes above 40 micro-Farads
(uF). Their voltage rating is usually 380-415 V, though sometimes 220-240 V types are also available.
Capacitor prices are approximately proportional to their voltage rating. Hence, higher voltage capacitors
are generally better value, since the volt-amp capacity increases as the square of the voltage.
'Motor run' type capacitors are generally quite cheap and, even for a star connection, should be less
than 30% of the generator cost. 'Motor start' capacitors must be avoided, since they are not designed
for continuous use. They are normally labelled 'motor start' or electrolytic and can also be recognized by
their very low cost and large capacitance rating for their size. Capacitors designed for power factor
correcting fluorescent lamps can be used with induction generators. However, they are rarely available
in sizes above 10 uF.
Capacitor life is governed by quality of manufacture and operating voltage, frequency and temperature.
The lifetime is very dependent upon voltage. For example, a capacitor that lasts a year at its rated
voltage may last for thirty years at half rated voltage. Therefore, it is recommended that 415 V
capacitors are used even for star connected arrangements, and that 415 V operation is avoided unless
the capacitors are rated for more than 415 V.
Capacitor life is also dependent upon frequency. An increase in frequency will increase capacitor
current, adding to the losses and thereby raising the temperature of the insulation.
Capacitor life is also highly dependent upon ambient temperature. The capacitors should be placed in a
box that has ventilation holes or
slots to help cool the capacitors, and this should be placed in a cool place away from direct sunlight.
Capacitors are only available in standard sizes and are generally specified with a tolerance of +/- 10%.
Hence, without measuring individual capacitors it is difficult to obtain the precise capacitance required.
The next section shows that great accuracy is not required, though efforts should be made to be within
10% of the value calculated.
Calculation of excitation capacitance
A precise calculation of the capacitance required to generate a given voltage under specific load
conditions is only possible with an accurate knowledge of the electrical parameters of the induction

machine, including the variation of the parameters with voltage. These parameters can be obtained by
means of a number of standard tests, but expensive equipment is required.
In practice it is sufficient to calculate an approximate value of excitation capacitance and adjust the
turbine speed until the required system voltage is obtained. This will mean that the operating frequency
may differ from the rated frequency of the induction machine, which is acceptable provided that the
frequency is kept within reasonable limits. These limits are explained in Chapter 5.
Two methods are recommended for the approximate calculation of the excitation capacitance required:
an electrical test method and a technique requiring only induction motor performance data.
No load test
The electrical test method involves a 'no load test' either as a motor or a generator. No load test results
can be used to calculate excitation capacitance because the apparent power drawn by the induction
machine at no load is approximately equal to the reactive power required when running as a generator
at close to full load.
The motoring test is simplest, but it requires a three-phase supply with sufficient capacity to start the
machine (the starting current will be approximately six times the rated full load current). A voltmeter
and clip-on ammeter are required to measure the line voltage and current, in order to determine the
apparent power and therefore the reactive power to be supplied by the excitation capacitors. A sample
calculation is given in Example 3.
Example 3
A 2.2 kW, 4 pole, 50 Hz, 415 V, 3 phase induction motor draws a current of 3.5 A when supplied at its
rated voltage and frequency and run as a motor with no mechanical load. Calculate the excitation
capacitance which must be connected in star to make the machine generate at approximately its rated
voltage when driven at slightly above its rated speed.
The total apparent power at no load,
SSnol0ad = ^3x Vlinexllme* = V3x415x 3.5 = 2516 VA
Using the relationship that the no load apparent power equals the reactive power to be provided by the
excitation capacitors (as given in the no load test section), the total reactive power,
IQ = 2:Snoload =2516VAR Hence, the reactive power per phase,
Q 2516 Qphase =- =------= 837VAR
33
For star connected capacitors,
Since,
Xc = =-------, C =------(6)
I 27tfC
27rfV
3.5
Hence,
C =-------------= 46 |oF
2ti.50.240 =
For derivation see Appendix 1.
Note that for a star connection the capacitor current is the same as the no load line current.

If there is no three-phase supply of sufficient capacity to start the induction machine, then the no load
test can be performed with the machine operating as a generator. A small single or three-phase motor is
used to belt drive the test machine. The test machine should be driven at very close to synchronous
speed and capacitors connected to excite the machine to rated voltage.
Provided the test machine is driven at very close to synchronous speed, this is more accurate than the
motoring test though it takes much longer to set up. It is generally only worthwhile for old machines
with no performance data.
The no load test method underestimates the capacitance required. However, it is quite accurate for
machines below 5 kW.
Manufacturers' data
The second method uses manufacturers' data to estimate the reactive power required from the
capacitors, by means of the full load current and power factor.
The capacitive reactance for full power factor correction can be obtained from this information and the
rated voltage. If the manufacturers' data is unavailable then an estimate can be obtained using the data
given in Table 1. This will be fairly accurate for machines of modern design. A sample calculation is given
in Example 4. Further practice can be obtained by repeating this calculation for the machine of Example
3 and comparing the results.
Usually, if either of these two methods for capacitance calculation are used, the generator will produce
too low a voltage when operated at rated frequency. The reason for this is that they do not fully
compensate for the increased saturation when the machine is operated as a generator under load.
However, the use of extra capacitance to allow for the increased saturation is not recommended. There
are two reasons for this:
1) A high level of saturation will reduce generator efficiency and maximum output, as shown in Example
5. There are two much more preferable alternatives: operation at reduced voltage and operation at
increased frequency, as explained in Chapter 5.
2) Manufacturers' data is not always reliable. If too much capacitance is connected then under
frequency operation will occur, which is inefficient for the generator and may damage loads. In general,
too little capacitance is better than too much.
Example 4
A 15 kW, 4 pole, 50 Hz, 415/240 V induction motor is star connected for operation as a generator. Using
the performance data given in Table 1, calculate the amount of excitation capacitance that should be
connected in delta to make the machine generate at approximately its rated voltage, when driven at
slightly above its rated speed.
From Table 1, the full load line current is 27 A and the power factor is 0.87. The total apparent power at
full load,
ZS = V3 x Vlmc x Ilme = V3 x415x27 = 19408VA
Hence, the real power,
2P = ZScos0 = 19408x0.87 = 16885W
The reactive power can be obtained from the power triangle:

It is a good idea to select combinations of capacitors that will allow some fine-tuning to be done on site.
This can be used to ensure that the operating frequency is acceptable and to maximise the power
output, as explained in Chapter 11.

5. OPERATING VOLTAGE AND FREQUENCY


The operating voltage and frequency of the generator will depend upon the turbine power output, the
excitation capacitance and the load connected, including its power factor. There is some flexibility as to
the choice of operating voltage and frequency which can be used to advantage to improve the efficiency
of the system.
Operating voltage
Generator considerations
As explained in Chapter 2, an induction generator can be operated over a range of voltages. Figure 4
showed how, for constant speed operation, the voltage is determined by the amount of capacitance
connected.
For operation at synchronous speed, the highest voltage that can be achieved is typically 125% of the
motor rating and the lowest voltage approximately 65% of the motor rating. The upper limit is set by the
voltage at which the excitation current equals the rated current of the machine and the lower limit is
the voltage at which the machine is sufficiently saturated to operate stably.
Operation at close to the upper voltage limit is not recommended, because, due to winding
temperature considerations, the large current flowing into the generator from the excitation capacitors
reduces the maximum load that can be connected and reduces the efficiency of the generator. This was
shown in Example 5. Operation at close to the lower voltage limit is not recommended for two reasons:
1) The reduced excitation capacitance makes the generator very sensitive to reactive loads, which
could cause significant variations in frequency and possibly loss of excitation, as explained in Chapter 6.
2) The efficiency will be below the optimum efficiency, particularly when the generator is heavily
loaded. This is because, in order to achieve the same power output at reduced voltage, the load current
must be increased. For example, a 33% reduction in voltage will require a 50% increase in load current
in order to maintain power output. This increases stator and rotor heating and reduces efficiency.
As shown in Appendix 2, to achieve optimum efficiency the induction machine should be operated as a
generator at 90-95% of the
motor rating. In addition to improving efficiency, this cuts capacitor costs and improves insulation life.
When selecting the operating voltage for the generator, transmission line voltage drop must also be
considered.
Load considerations
Voltages above the manufacturers' ratings should be avoided as the lifetime of the equipment, be it a
heater, lamp or motor will be reduced. Manufacturers' data indicate that for a standard incandescent
lamp a continuous overvoltage of just 5% will decrease the expected life of the lamp by approximately
50%. Hence, it is clear that operation of equipment at above rated voltage should be avoided.
The effect of an undervoltage with a heater or incandescent lamp is to increase the life and decrease
the output. In the case of fluorescent lamps too low a voltage can prevent turn on and may cause the
starter to fail due to repetitive operation. With motors, the starting torque will be reduced and
overheating may occur if the motor is fully loaded.
Electrical equipment is normally designed to work for a supply voltage variation of+/- 6%, and the lower
voltage limit can be extended to -15% for systems where just resistive loads and fluorescent lamps are
used. Below this voltage incandescent lamps will operate very inefficiently and starting problems will
occur with fluorescent lamps.
Operating frequency
Generator considerations
Increasing the operating frequency will reduce the excitation current required to achieve rated voltage,
as shown in Figure 12. This results in improvements in efficiency and maximum power output, as shown
in Appendix 2.
Load considerations
The operation of some domestic electrical appliances, such as television sets, record players and clocks
was, at one time, dependent upon their synchronization to the supply frequency, and they therefore
required a well regulated supply. This is not the case with more modern equipment since an internally

generated reference frequency is used. Nowadays, the limits of acceptable frequency variation are
principally set by the requirements of transformers and motors.

Figure 12 No load excitation characteristics for an machine operated at two different frequencies
induction
The effects of supply frequency on motor performance depends on the type of motor used. Only two
types of motor are commonly used on a.c. supplies; these are universal motors and induction motors.
The universal motor is so called because it gives comparable performance on both d.c. and a.c. supplies.
Conversely, induction motor performance is greatly affected by the supply frequency and as a result
these machines provide the main limits to its value.
The magnetizing current for an induction motor increases if the supply frequency falls, in the same way
as for an induction generator. This increases the power dissipation in the stator windings and can cause
them to overheat. This same effect occurs with transformers. Small transformers, which are often used
in electrical appliances, are very susceptible to damage from such under frequency operation. For these
reasons, operation at below rated frequency should be avoided.
Increasing the operating frequency has the opposite affect. It reduces the magnetizing current and
therefore induction motors and
transformers run cooler. However, there are disadvantages. The main constraint occurs if motor loads,
such as pumps and fans, with a power requirement that increases significantly with speed are used on
the supply. The effect of such loads can be appreciable since, for induction motors, shaft speed
increases approximately linearly with frequency, and the power requirement of fans increases at nearly
the cube of the speed. A 10% increase in frequency is generally acceptable, even with such highly speed
dependent loads. The reasons for this are as follows:
1) Because of oversizing, most motors run at only 60 - 80% of their rated output when operated at
rated frequency. Hence, there is some spare capacity.
2) As previously explained, higher frequency operation reduces the magnetizing current, thereby
offsetting some of the increase in load current.
3) Increased shaft speed increases the cooling of the machine.
If no highly speed dependent motors are used then the upper frequency limit can be increased to 20%
above rated frequency. If only resistive loads are used then there is no frequency limit imposed by the
load. The upper limit will then be determined by the limit of stable operation of the generator, i.e. the
frequency at which the machine is no longer saturated for the required voltage output.
Using the operating voltage and frequency to advantage
If an induction machine is operated as a generator at the same voltage and frequency that it is designed
for as a motor it will run at lower efficiency and higher temperature than for an equivalent power
output as a motor. This is due to increased saturation and therefore greater magnetizing current.
Careful selection of the rated voltage of the induction machine and the operating voltage and frequency
can reduce the level of saturation and result in the following advantages:
Improved generator efficiency

Increased maximum generator power output (within thermal limits)


Increased winding life
reduced capacitor costs
The operating voltage is determined by the voltage rating and voltage requirements of the load and the
transmission line voltage drop. Ideally, the induction machine should be operated as a generator at 6%
below the motor voltage rating, to give the advantages listed above. For example, to operate at 240 V a
255 V induction motor should be used. Chapter 3 presented a number of options for obtaining a
machine with the correct voltage rating. If the desired motor voltage rating cannot be obtained then
consider operating the loads at slightly above their rated frequency, as this will also provide the
advantages listed above.
Though less efficient, it is quite acceptable to operate the generator at the same voltage and frequency
that it is designed for as a motor. More capacitance will be required, compared to the values calculated
using the methods given in Chapter 4, and extra care should be taken to ensure that the winding
currents are not too high.
6. THE EFFECT OF LOAD UPON GENERATOR OUTPUT
Resistive load
Figure 13 shows the voltage variation with resistive load for an induction generator operating at
constant speed and with fixed excitation capacitance. The full load current is the current at which the
generator current equals the rated current as a motor. Excessive load will cause excitation to collapse.

Figure 13 Typical voltage against load current for an induction generator with fixed excitation
capacitance operated at constant speed
Figure 13 does not provide a full representation of what occurs when the generator is used on a micro
hydro scheme, since it takes no account of frequency variation due to the power-speed characteristic of
the turbine. A typical turbine power speed characteristic is shown in Figure 14. When the load on the
generator is increased the voltage will drop immediately by an amount determined by the characteristic
of
Figure 13. The increased load on the turbine will cause the turbine speed to drop, resulting in a
corresponding reduction in operating frequency. This drop in frequency causes a further drop in voltage,
due to reduced excitation, as shown in Figure 12. The turbine speed and generated voltage will fall until
a speed is reached at which the power output of the turbine equals the load on the turbine.

Conversely, if the resistive load on the generator is reduced the generated voltage will rise and the
turbine speed will increase. If all the load on the generator is disconnected the voltage and turbine
speed will increase until the losses in the generator equal the power output from the turbine.
It is clear from the above that the resistive load must be controlled in order to regulate the voltage and
frequency. Load control options are presented in Chapter 9.
Inductive load
If an inductor is connected across the output of the generator the current into the inductor will be
supplied by the capacitors connected to the generator. This will reduce the amount of excitation
capacitance available, and as a result the voltage will fall temporarily. The drop in voltage will reduce
the power dissipated in the other loads and therefore the turbine speed will increase. The voltage will
increase as the speed increases until equilibrium is reached at a speed where the power output of the
turbine equals the load on the turbine. Hence, inductive loads result in an increase in operating
frequency.
In an uncontrolled system the new operating voltage will depend largely upon the power-speed
characteristic of the turbine. If the turbine power output does not vary appreciably between the initial
and final speeds there will be no significant change in voltage.
If too large an inductive load is connected the excitation of the generator will collapse. This can occur
when connecting induction motors as loads, as discussed in Chapter 10.
Capacitive load
In the same way that an inductive load will increase the operating frequency, a capacitive load will
reduce it. This could be damaging to the generator and other loads, as explained in Chapter 5.
Fortunately, leading power factor loads are very rare. They may occur if too much power factor
correction capacitance is connected across an inductive load.
7. SINGLE-PHASE OUTPUT FROM A THREE-PHASE MACHINE
Single-phase induction motors can be used as generators, but problems can be experienced in achieving
excitation and in determining the size and arrangement of the capacitors required. In addition, singlephase induction motors are more expensive than three-phase induction motors and are only available
for small power outputs. Fortunately, it is possible to use a three-phase induction motor as a singlephase generator and this is the preferred approach to providing a single-phase supply. The method for
obtaining a single-phase output from a three-phase machine is as follows:
1) Use a three-phase induction machine suitable for 240/415 V operation and connect the machine in
delta.
2) Calculate the per phase capacitance, 'C, required for normal three-phase 240 V delta operation.
3) Instead of connecting 'C to each phase connect twice 'C to one phase, 'C to a second phase and no
capacitance to the third phase. This is known as the 'C-2C' connection. The load should be connected
across the 'C phase, as shown in Figure 15.

Figure 15 Single phase generation from a three phase machine using the 'C-2C connection
This unbalanced arrangement of capacitance helps to compensate for the unbalanced load on the
generator, and as a result the generator can be used with an output of up to 80% of the motor rating.
This is the same derating factor that is applied to three-phase machines to compensate for load
imbalance.
In using the 'C-2C connection it is essential to ensure that the direction of rotation of the machine rotor
is correct, in relation to the phases to which the 'C and '2C values of capacitance are connected. If the
capacitors are arranged correctly the 'C phase' will produce its peak voltage before the '2C phase'. If the
opposite occurs the generator will run inefficiently and overheat.
It is possible to determine the correct rotation from the labelling of the windings. However, it is quite
easy to make mistakes or to find that the windings have been incorrectly labelled. Correct rotation
should be checked by running the machine with both phase sequences. This is a relatively
straightforward task. The correct rotation can be determined by measuring the power output or the
winding currents.
Power output test
For this test a constant input power is required. This is usually achieved by running the turbine at
maximum power output. The turbine output is gradually increased and load applied to the 'C phase' to
maintain the voltage at its rated value. When maximum turbine power is reached, measure the load
current. Repeat the test with the '2C capacitance connected to the phase that previously had no
capacitance across it. A higher load current reading will be obtained for the correct capacitance
arrangement because the generator will run more efficiently. A 10% increase in electrical output is
typical. The efficiency improvement will mean that much less power will be dissipated in the generator
and as a result it will run appreciably cooler.
This test is very easy to perform if an induction generator controller is used because the controller will
maintain a constant voltage and the ballast meter will indicate which connection produces the
maximum power output.
Winding currents test
By detailed analysis, it can be shown that for a purely resistive load, if the condition given in Equation 7
is met, the 'C-2C connected generator behaves as a balanced three-phase machine. Hence, if the
8. GENERATOR PROTECTION
Overload
Provided that the generator rating is greater than or equal to the maximum electrical output from the
turbine-generator, an overload of the generator windings, due to excess load on all phases, will not be
possible. The reason for this is that the partial collapse of excitation caused by the overload will limit the
winding currents to safe values. However, with a three-phase system a severe overload on one phase
could cause that phase to be burnt out. Miniature circuit breakers (MCBs) should be used to prevent
this. MCBs rated for 1.5 times the rated full load generator output can be used, since the machine
output has been derated to allow for some load imbalance. If three-phase motor loads are being used
then a three-pole MCB should be used to prevent single phasing of the motor if the trip operates.
If the generator rating is less than the maximum electrical output from the turbine-generator, MCBs
should be fitted to the generator output to prevent damage due to overloading. Note that irrespective

of whether a three-phase or single-phase 'C-2C system is used the maximum generator rating is 80% of
the motor rating.
Underload
The induction generator is at greater risk from loss of load than from too much load. The reason for this
is that, with little or no load connected, the generator will speed up and hence the frequency and
voltage will increase. Both of these factors increase the current flowing into the generator from the
capacitors and hence the winding current will increase, generally to a level that exceeds its rated value.
To protect against damage to the capacitors and the windings, MCBs should be fitted in series with the
capacitors in order to switch them out if the generator overspeeds. Such an arrangement is shown in
Figure 16. The current rating of the MCBs should be slightly above the maximum capacitor current
under normal operating conditions.
It is preferable to disconnect all the capacitors at once. Hence, a three-pole MCB should be used with a
three-phase system and a two-pole MCB with a 'C-2C single-phase system. The two-pole MCB should be
rated for the maximum current through the '2C set of capacitors. Single-pole MCBs can be used if two
and three-pole versions are unavailable.

Figure 16 Underload protection by means of MCBs.


Magnetic MCBs are preferable to thermal MCBs as they are more accurate and unaffected by
temperature variations. The MCBs should be rated for between 1.2 and 1.5 times the capacitor current
under normal conditions. If the MCB has too high a current rating then there is a danger that it will not
trip when all the load is removed. Standard current ratings for MCBs are 2, 4, 6, 10, 16, 20, 25, 32, 40
and 50 Amps.
The current setting of some MCBs can be manually adjusted. If MCBs with a suitable current rating
cannot be obtained then a combination must be used. For example, if a three-phase generating system
has a capacitor current at rated frequency and voltage of 10 Amps per phase the MCBs should be rated
for between 12 and 15 Amps. Since there is no suitable MCB within this range the capacitors in each
phase should be split into two equal sets and protected by 6 Amp MCBs.
The correct operation of the MCBs should be checked by performing a test at runaway speed when the
generator is installed. This is done by disconnecting the consumer load and any ballast load, and
opening the turbine valve fully. The power output of the turbine will be dissipated in the generator and
drive. If the MCBs fail to trip
within ten minutes then they must be replaced with MCBs of a lower current rating or greater
sensitivity. For small overcurrents MCBs take several minutes to operate.
Small generating systems tend to have the lowest runaway voltage, because, as shown in Table 1, the
induction machine is less efficient and will therefore dissipate the turbine power output at a lower
voltage. The runaway voltage is typically twice rated voltage, though for small generators it can be as
low as 360 V and for large generators as high as 600 V.
If the turbine operating at runaway speed does not cause the excitation current to exceed the rated
current of the generator then MCBs are not necessary, provided that the capacitor voltage rating is not
exceeded. This is only likely for turbines with an efficiency that falls off rapidly with overspeed.
If an induction generator controller is used the small amount of capacitance that this contains could be
sufficient to cause excitation to persist, even once the main capacitors have been disconnected by the

MCBs. This is acceptable since the excitation current under these conditions will be considerably below
the rated current of the generator.
Normally when the turbine is closed down the capacitors will discharge into the generator windings.
However, if the capacitors are disconnected by the MCBs they will remain charged and could cause an
electric shock if touched. The resistors shown in Figure 16 are incorporated for electrical safety as they
will ensure that the capacitors are discharged when the turbine is closed down. For the sake of high
reliability, ensure that under normal operating conditions the power dissipation in the discharge
resistors is less than 25% of their rated power dissipation.
Overspeed
Standard induction machines can withstand considerable overspeeds due to their solid rotors. Twice
rated speed is normally quite acceptable, though for 2 pole machines and large machines
manufacturer's advice should be obtained. Most micro-hydro schemes use 4 pole generators, with
turbines having a runaway speed less than twice rated speed, and therefore the induction generator
needs no overspeed protection.
9. FIXED AND VARIABLE LOAD SYSTEMS
Fixed load systems
The simplest induction generator systems have a constant turbine power output and a constant load.
No form of load control is required and therefore the cost and complexity of the system is kept to a
minimum.
The system is operated by bringing the turbine up to speed and then switching in the electrical load. The
turbine power output is then increased until the required voltage is achieved. An overvoltage trip should
be used to prevent loads from being damaged if some of the electrical load becomes disconnected. The
design for a suitable trip is given in Appendix 5.
The fixed load approach has a number of disadvantages. These are as follows:1) It is often difficult to ensure a constant electrical load. For example, on a village micro hydro scheme
it might be assumed that a load consisting of light bulbs only, with no individual switches, represents an
ideal fixed load. Experience in Nepal has shown that this is not always the case, as villagers may remove
the bulbs from their sockets if they decide to go to bed early.
2) Tripping of the system due to overvoltages can be very inconvenient, especially if the generator is far
from the load centre. As a result the normal operating voltage may be set at a low level to reduce the
chances of tripping. This results in inefficient operation of some domestic appliances, particularly light
bulbs.
3) Frequent tripping of the system may result in the overvoltage trip being bypassed by the users.
4) Switchable or variable loads must be avoided. If used they must only make up a small proportion of
the total load. For example, a refrigerator can generally be used on a system where there is a large fixed
load.
5) In general, less productive use is made of the power available.
To maximise the use of fixed load systems, changeover switches can be installed allowing a choice
between two or more appliances of approximately the same power rating. However, a system with an
overvoltage trip and changeover switches is less versatile than one with a load controller and may be
almost as expensive.
Variable load systems
For a small increase in overall system costs, an electronic controller can be used with the generating
system to allow large changes in consumer load to be accommodated. This overcomes the
disadvantages for fixed load systems that were listed in the previous section.
The generated voltage and frequency are controlled by maintaining a near constant load on the turbine,
as shown in Figure 17. The controller compensates for variations in the main load by automatically
varying the amount of power dissipated in a resistive load, generally known as the "ballast load', in
order to keep the total load constant.

Figure 17 Basic principles of electronic load control


The first electronic load controllers were designed for use with synchronous generators. They sense and
regulate the generated frequency. Voltage control is not required because synchronous generators have
inbuilt voltage regulators. Without an electronic frequency controller, the frequency will vary as the
load changes and under no load conditions will be much higher than rated frequency.
The control requirements of the induction generator are different from those of a synchronous
generator, since, without any electronic control, both the voltage and frequency will vary when the load
changes. As a result, early control systems contained a voltage
controller as well as a frequency controller. The voltage controller controlled excitation by varying the
amount of capacitance connected and the frequency controller varied the resistive load, just like the
controllers for synchronous generators. Such systems are expensive and complex and therefore have
not been widely used.
The induction generator controller (IGC), developed by the author for ITDG, dispenses with the need for
a separate voltage controller by using the characteristics of the generator and water turbine to
advantage. The controller is very similar to the electronic controllers used with synchronous generators,
except that voltage is sensed and directly controlled instead of frequency. The characteristics of the
generator are such that this controller also produces good frequency regulation.
When a resistive load is connected to the generator the voltage falls. The IGC senses the drop in voltage
and compensates by decreasing the power dissipated in the ballast load. A small transient change in
frequency will occur before both the frequency and voltage return to their original values. The total load
on the generator will be the same as before the load change.
If an inductive load is connected to the supply the voltage will fall, and hence the controller will reduce
the power dissipated in the ballast. This reduction in resistive loading causes the turbine to speed up,
increasing the generated frequency. The increased frequency reduces the magnetizing current required
by the generator and increases the capacitor current. As a result, the voltage will rise and return to its
initial value. The frequency will have increased in order to compensate for the inductive load.
As a result of the high level of saturation of modern induction machines the frequency increase with
inductive loads is small, typically less than 10% for an overall load power factor of 0.9. Given that the
ballast load is resistive, it is not difficult to maintain an overall load power factor greater than 0.9.
Improved frequency regulation can be obtained by power factor correcting significant inductive loads. If
a large number of fluorescent lamps are used on the system this will constitute a significant inductive
load.
Controller options
There are a number of options for varying the ballast load: phase angle control, binary weighted loads
and mark-space ratio control.
Phase angle control

Figure 18 Ballast current waveform for a phase angle controller


By delaying the switching on of a triac or thyristor arrangement a variable resistive load can be
produced, as shown in Figure 18. Each half cycle, the switching on of the ballast load is delayed by the
phase angle a which can have a value of between 0 and 180.
This control approach is often used with synchronous generators, but is less appropriate for induction
generators because of the variable lagging power factor produced due to the ballast current lagging
behind the voltage. The effect of this with an induction generator is to increase the frequency variation
that occurs, as explained in Chapter 6. A further disadvantage is waveform distortion which produces
increased heating in the generator windings. To compensate for the waveform distortion, the generator
should be oversized.
Binary weighted loads
In this case the variable resistive load is produced by switching in a combination of fixed resistors. The
values of these fixed resistors are binary weighted so as to achieve the maximum number of load steps
with the minimum number of resistors and switches. A single-phase, three resistor arrangement is
shown in Figure 19, along with the seven values of ballast load that can be achieved.

Figure 19 Binary-weighted load controller and ballast current range


The main advantages of this approach are that waveform distortion is not produced and the ballast load
is resistive. Binary weighted controllers have been used in Nepal. However, they do have a number of
disadvantages. The main disadvantage is the complexity resulting from requiring a number of ballast
loads, each with its connections, wires and switching device.
The ballasts must be of accurate resistance value, which may prove difficult with available heating
elements, especially for small power ratings. In addition, because the ballast load is only varied by fixed
steps, the voltage is only controlled within a range or 'window'. For stable operation under part flow
conditions a large window must be used, resulting in poor voltage regulation.
Mark-space ratio controller

Figure 20 Basic switching circuit for a single-phase mark-space ratio controller


The mark-space ratio controller, in its simplest form, requires just a single ballast load, as shown in
Figure 20. The ballast load is connected across the rectified output of the generator and switched on
and off by means of a transistor. The controller varies the on (mark) and off (space) times, known as the
mark-space ratio, in order to control the power dissipation in the ballast load. The on time can be varied
over the full range of 0% to 100%. Figure 21 shows voltage waveforms with 60% on-time.

Figure 21 Typical ballast voltage waveform for a single-phase mark-space ratio controller with 60% on
time

The advantages of this type of controller are good voltage regulation, simple connection of ballast loads
and an effectively resistive ballast load. It is the preferred IGC design and has now been installed on a
total of more than forty sites in Indonesia, Nepal, Sri Lanka and England. Figure 22 shows an IGC locally
assembled and installed in Nepal.
The three phase version uses a three phase bridge rectifier as shown in Figure 23. Its disadvantages are
that there is no phase balancing and there is increased waveform distortion. The advantages are that it
is as simple as the single phase design and uses the same circuit board.

Figure 23 Three phase mark-space ratio controller


The IGC has been developed for local manufacture in developing countries. A good knowledge of
electronics is required in order to manufacture, install and repair it. To ensure good quality local
manufacture the IGC design is only provided to engineers who attend an ITDG approved training course.
10. MOTOR STARTING

Motor starting can be a considerable problem with induction generators. The reasons for this are that
the starting current for an induction motor is typically 6 times the rated current and the power factor at
starting is lower than when the motor is running. For direct-on-line starting the limit for the motor, as a
proportion of the generator rating, is typically 15% for a single-phase motor and 10% for a three-phase
motor.
Single-phase motors are generally easier to start because their starting current is not as high and their
power factor is better. If too large a motor is switched on there will be insufficient capacitance to both
supply the generator and power factor correct the motor, and the generator will de-excite.
These values are just guidelines, since a lot depends on the load that the motor is driving. Loads
requiring a high starting torque, such as compressors, are much harder to start than low starting torque
loads such as fans. For example, a 1.1 kW single-phase fan could be started by a 4 kW induction machine
connected 'C-2C and producing 2 kW. Whereas, all kW induction machine connected 'C-2C' and
producing 800 W was unable to start a 180 W refrigerator compressor.
Techniques for improving motor starting
One option for improving motor starting is to oversize the generator, since this provides extra
capacitance that will help to power factor correct the motor at starting. The disadvantages of this
approach are increased cost and reduced efficiency, especially at part load. Better options are power
factor correction at starting or, for three-phase machines, star-delta starting.
Power factor correction at starting involves temporarily connecting capacitors to compensate for the
inductive current. The starting current can usually be obtained from manufacturers' data sheets. It is
generally called the locked rotor current, and refers to the initial starting condition when the rotor is
stationary. The power factor at starting is low, typically 0.4. Hence, at starting the inductive current is
only slightly less than the locked rotor current.
Example 8
For the 5.5 kW motor described in Example 7 the locked rotor power factor is 0.4. Calculate:
(a) The power drawn from the supply for a direct on line start.
(b) The additional capacitance required for full power factor correction at starting, assuming that the
motor is already compensated for full load running.

Subtracting the 27F connected for full load power factor correction, the additional capacitance
required is,
The additional capacitance required at starting is more than ten times the capacitance required when
running.
The capacitance required to power factor correct at starting is much higher than that required when
running, as shown by Examples 7 and 8. If standard capacitors are used then the cost of the power
factor correction will be very high. However, since the capacitors need only be connected for a few
seconds, whilst the motor runs up to speed, capacitors rated for intermittent operation can be used.

'Motor start' electrolytic capacitors are the obvious choice since they are designed to be operated for
the short time required for motor starting. Their cost per micro-Farad is typically between 10% and 20%
of the cost of 'motor run' capacitors.
For switching the capacitors it is best to use solid-state relays (SSRs) of the type that switch at zero
current to reduce voltage transients. Contactors are not recommended because the arcing that occurs
at switching will rapidly wear them out. A control circuit for a single-phase system is given in Appendix
5.
With three-phase systems star-delta starters are a cheaper and simpler option provided that the motor
is rated 415 V when delta connected and the load is such that the motor will start with 1/3 of the directon-line starting torque.
With star-delta starting the motor is initially connected in star and once up to speed it is reconnected in
delta. This is done manually or automatically depending on the type of starter used. The two winding
configurations are shown in Figure 24.

Figure 24 Star and delta starting conditions


At starting, the per phase impedance, Z, is the same whether the motor is connected in star or delta.
Hence,
Ia=V3^ =VP^
V3.(V3.Is.z) Hence, IA =-------------------- =3.IS
(9)
Since the line current is reduced by a factor of three for a star connection, The real power and reactive
power will be reduced by a factor of three at starting.
Example 9 shows that by using a star-delta starter the starting capacitance can be reduced by a factor of
more than three. As a result, it is usually possible to start motors with a rating of one third of the
capacity of the generator. For example, at an installation in Indonesia with an 18.5 kW induction motor
running as a generator, and an output of 10 kW, it was possible to start a saw mill motor of 7.5 kW. To
achieve this a star-delta starter was used and the motor was power factor corrected for the running
condition.
Example 9
Repeat Example 8(b) for star-delta starting.
The reactive power is reduced by a factor of three. Hence,

11. GENERATOR COMMISSIONING


Safety
Wiring and earthing should be carried out to national standards and tested thoroughly. Mechanical
safety, including provision of adequate guards, should be checked before starting the turbine.
Output frequency and power
When the generator is commissioned, adjust the turbine power output in order to obtain rated voltage.
Then, whilst maintaining a constant voltage, vary the load between no load and maximum generator

output and measure the range of generated frequency. Check that the frequency range is acceptable
(see Chapter 5 for details) and if not change the amount of excitation capacitance connected. If
inductive loads are to be used these must be connected in order to determine whether the increase in
generated frequency is acceptable.
Check the winding currents at maximum load to ensure that the current rating of the induction machine
is not exceeded. Due to inbalance with the 'C-2C connection it is acceptable for one of the windings to
be overloaded, provided that the other two are underloaded. The following condition must be met:
If+I^+I^ 3x4ed
(10)
Where It ,I2,13 are the three measured winding currents Irated is the rated winding current
If the generator rating is less than the maximum electrical output from the turbine generator, either fit a
larger generator or ensure that there is adequate overload protection, as explained in Chapter 8.
Phase rotation
If a three-phase system is installed the phase sequence should be checked to ensure correct rotation of
any three-phase motor loads that are used.
If a single-phase 'C-2C system is used it is essential that the correct phase sequence is used to prevent
the generator from overheating, as explained in Chapter 7.
Underload testing
The testing of underload protection should be carried out, as explained in Chapter 8.
Adjustment for maximum efficiency
The system efficiency can sometimes be improved by adjustment of the amount of capacitance
connected. The resulting change in speed will cause a change in both turbine and generator efficiency
and, in the case of a reaction turbine, the penstock losses. The frequency must be kept within the
acceptable limits given in Chapter 5.
APPENDIX 1: Three-phase circuit theory
This section presents a summary of the essential terms and relationships used in the main text, along
with worked examples. For more comprehensive coverage refer to electrical engineering textbooks that
cover three-phase circuits.
Line and phase quantities
The terms line and phase are applied to voltages and currents when describing three-phase circuits, and
therefore it is important to be clear about their meaning.
In a single-phase circuit, the two wires connecting the load are called the line. However, if one of the
wires is earthed, the unearthed wire is usually referred to as the line to distinguish it from the earthed
one, which is called the neutral. In a three-phase four-wire system the line consists of four wires; three
lines (Lp L2, L3) and the neutral (N), as shown in Figure A 1(a). The voltage between any two lines is
referred to as the line-to-line voltage, but often shortened to the line voltage. If the load is balanced
(equal on each phase) the current in the neutral conductor will be zero and this conductor can be
dispensed with, thereby making a three-wire system. An alternative three-phase three-wire system,
again consisting of just three lines (L1;L2,L3) is shown in Figure A 1(b).
Figure A1 Voltage and current relationships in three-phase circuits

A phase is one of the three branch circuits making up a three-phase circuit. In a star connection, a phase
consists of those circuit elements connected between one line and neutral. In a delta connected circuit,

a phase consists of those circuit elements connected between two lines. Hence, in a star circuit the
phase voltage is the voltage between line and neutral and in a delta circuit it is the voltage between two
lines. The phase current is the current that flows through the circuit elements that make up the phase.
Three-phase supplies and loads are specified in terms of their line-to-line voltage and line current.
Hence, motor data, such as that given in Table 1, is specified using line values. Since this is the standard
representation for three-phase supplies, there may be no explicit reference to indicate that line values
are being used.
Voltage and current relationships
It is clear from Figure Al that for a star connected system the line currents (I,,^) equal the phase currents
(Ia,Ib,Ic), and that for a delta connected system the line voltages equal the phase voltages. By means of
phasor diagrams, it can be shown that for a star connected system the line voltage is V3 times the phase
voltage and for a delta connected system the line current is V3 times the phase current. These
relationships are summarised in Table Al.
Quantity Star connection Delta connection
Voltage V =V3 Vline phase V = Vline phase
Current line phase
I =V3 Iline phase
Table A1 Voltage and current relationships in three-phase circuits
Power and Power factor
In a.c. circuits the product of the r.m.s. values of the applied voltage and current supplying a load is the
apparent power VI voltamperes, the latter term being used to distinguish this quantity from the real
power, expressed in watts. For a resistive load the real power equals the apparent power. For a partially
inductive or capacitive load the real power is less than the apparent power, and the latter has to be
multiplied by a quantity termed power factor to give the power in watts.

Figure A2 Power triangle


A power triangle can be drawn, as shown in Figure A2. The third side of the triangle represents the
reactive power. The relationship between apparent, real and reactive power is:

For a delta connected load the result is the same, since:


St0ui=Vx^x3 =V3VI
V3~
Hence, for calculating the power drawn by a motor, by means of manufacturers' data, it is not necessary
to know whether it is star or delta connected.
Example A1
A 3 kW, three-phase motor, connected to a 415 V 50 Hz supply, draws a line current of 6.5 A. If the
power factor is 0.8, calculate:
(a) The apparent power supplied.
(b) The real power supplied.
(c) The reactive power supplied.
(d) The efficiency, assuming that the power output of the motor is 3 kW.
(a) The apparent power is:

S = V3VI =V3x415x 6.5 = 4672 VA


(b) The real power is:
P = Scos0 = 4672 x 0.8 = 3738 W
(c) The reactive power can be obtained from the power triangle:

For stand-alone induction generators, capacitors must be used to fully power factor correct the
induction machine, i.e. increase the power factor to 1.0, by providing sufficient VARs to meet the VAR
requirement of the machine. Example A2 illustrates the calculations required.
Example A2
For the motor specified in Example Al, calculate the capacitance required for full load power factor
correction:
(a) For star connected capacitors.
(b) Delta connected capacitors.
As calculated in Example Al, the total reactive power required by the motor , Q, is 2803 VAR.
Hence, the reactive power per phase,
Q 2803 Qphase =T = = 934 VAR

(b) DELTA: Vphase = Vlme =415V

APPENDIX 2: Induction generator efficiency


The following tables and graphs present efficiency and winding temperature results for a 230 V, 50 Hz,
three-phase, 4-pole, 2.2 kW induction motor when operated as a motor and a generator. For generator
operation the excitation capacitance was varied as the load was changed in order to maintain a constant
voltage at fixed frequency.
Load
Temperature Efficiency
(% motor rating) (C)
%
25
60
71.8
50
65
81.6

75
73
84.1
100
82
83.4
125
91
81.6
Table A2 Motor performance (50 Hz and 230/400 V star connected)
Connection Load(% motor rating) Capacitance Temperature (C) Efficiency %
Balanced Delta 25 50 75 100
47 52 58 66 64 76 82 92
65.674.2 75.8 75.9
C-2C
25 50 75 100
43 50 58 67 66 75 86 95
64.3 73.9 75.8 75.6
Table A3 Generator performance for balanced delta and 'C-2C connections (50 Hz and 230V)
Frequency
Load(% motor
Capacitance
Phase Voltage
Temperature(C) Efficiency %
(Hz)
rating]
(uF)
230 230 230 230 50 50 50 50 25 50 75 100
44 48 55 63
59 69 81 91
67.3 75.4 78.1 78.0
215 215 215 215 50 50 50 50 25 50 75 100
37 4047 55
46 54 68 82
71.1 78.0 82.3 79.8
200 200 200 200 50 50 50 50 25 50 75 100
33 37 43 50
47 56 68 80
70.6 77.6 80.5 78.8
230 230 230 230 55 55 55 55
72.8 80.1 81.5 81.1
25 50 75 100 125 28 31 35 40 46 52 60 68 76 86
230
55
80.4
Table A4 Generator performance for a star connected machine under a range of voltage and
frequency conditions.
The tables and graphs clearly show that induction machines are more efficient when operated as
motors than as generators. Improvements in generator mode performance can be obtained by
operating at less than rated voltage and more than rated frequency. For motor and generator mode
operation, maximum efficiency is generally achieved at about 80% of motor rating. Hence, restricting
maximum generator mode power output to 80% of motor rating, as well as protecting against excessive
winding temperatures, tends to improve efficiency.
With 'C-2C operation efficiency is always lower than for balanced three-phase delta operation, except at
the balance point. However, the difference in efficiency away from the balance point is quite small.
Delta operation is less efficient than star operation due to increased harmonic currents.
Figure A3 Efficiency curves for generator and motor operation (star connected and 50 Hz)

Figure A4 Efficiency curves for generator operation with various configurations (230 Vand 50 Hz)

Figure A5 Efficiency curves at 230 V and two frequencies

APPENDIX 3: Single transformer systems


If a long transmission line is required, it is worth stepping the voltage up and down by means of
transformers in order to reduce cable costs. However, the savings in terms of cable cost must be
compared with the expense and additional losses introduced by the transformers.
An intermediate option is to generate at a high voltage so that only one transformer, the step down
transformer, is required. This is particularly useful for single-phase systems, since all standard induction
motors can be connected for 415 V line to line and the 'C-2C connection applied.
415 to 240 V step-down transformers are relatively easy to obtain. Transmitting at 415 V rather than
240 V means that the cable cross-sectional area can be reduced by a factor of three for the same power
loss. The transformer represents a significant inductive load and should therefore be power factor
corrected.
APPENDIX 4: Halving the operating voltage of an induction machine
In most cases, each phase of a standard induction machine consists of an even number of groups of coils
connected in series. The voltage rating can be halved and the current rating doubled by reconnecting
the groups into two parallel sets. This is achieved by breaking and then rejoining the connections
between the groups of coils that are found in the end windings of the machine. No rewinding is
required.
This operation can be carried out by a competent machine winder or any engineer equipped with the
necessary heat resistant sleeving, lacing, varnish and brazing equipment. Ideally the windings should be
joined by brazing, as this produces a strong connection that is able to withstand vibration. If brazing
equipment is not available then the joints can be soldered using a heavy duty soldering iron. With
soldered joints extra care should be taken when varnishing and relacing to fix the joints so that they
cannot flex.
A series to parallel conversion for a machine with two groups of coils per phase will be described, since
this is commonly found with both two and four pole machines. The procedure is as follows:
1) Isolate all six ends of the windings at the terminal block.

2) Expose the connections in the end windings. These are usually at the drive end of the machine and
will be covered with heat resistant sleeving.
3) Pull the connections away from the windings, cutting the lacing as required.
4) Cut away the heat resistant sleeving to expose the connections.
5) For a machine with two groups of coils per phase there will usually be three brazed joints, per phase,
as shown in Figure A6(a). Leads from the terminal block connect to the ends of the groups (joints A and
C in the diagram). The middle connection (joint B) is between the two groups of coils. The three joints
can easily be identified using a multimeter with a resistance setting.
6) Label the wires corresponding to Ul and U2 in Figure A6(a). Cut joint B and using the meter identify
and label Ul' and U2'.

(a)
(b)
Figure A6 Series to parallel conversion of coil connections
7) Connect Ul* to U2 and U2' to Ul, as shown in Figure A6(b), covering the connections with heat
resistant sleeving. Note that the cables going to the terminal block will now carry twice their original
current. If these cables are not sufficient for the higher current then they must be replaced.
8) Repeat steps 5, 6 and 7 for the other two phases.
9) Relace the windings and joints and apply varnish to the heat resistant sleeving and any areas where
the old varnish has been disturbed.
10) Test the windings using an insulation tester.
If there are four or more groups of coils per phase, the middle connection (joint B) is at the middle of
the series. The same procedure can be used.
APPENDIX 5: Overvoltage trip
A circuit to protect the electrical loads in the event of an overvoltage is shown in Figure A7. This should
be used with all fixed load systems, as explained in Chapter 10. When an overvoltage occurs, the relay
contacts open and the contactor coil becomes de-energized causing its contacts to open and disconnect
the loads from the generator.
The only component that is dependent upon the generator rating is the contactor, all the others are
standard. The current rating of the contactor must be sufficient to supply the maximum load on the
generator. Two current ratings are usually given for a contactor: an AC1 rating and an AC3 rating. The
AC1 rating is for largely resistive loads and the AC3 rating for motor loads. Although, for a fixed load
system, the load will fall into the AC1 category, the AC3 rating should be used in order to ensure long
life.
Contactors usually have three main contacts (or poles as they are often called) and one auxiliary contact
with a lower current rating. For single-phase generation the main contactor poles can be connected in
parallel so that they share the current. A multiplying factor can then be applied to the three-phase
current rating. For two poles in parallel the factor is 1.6 and for three poles in parallel 2.2. This allows for
unbalanced current distribution between the poles.
Care should be taken to ensure that the generated voltage and frequency, under normal conditions, are
suitable for the contactor coil. If the voltage is higher and/or the frequency lower than the coil rating
then its life will be reduced. Reduced voltage and/or increased frequency operation is normally quite
acceptable provided that the combined effect is less than 25%. For example, if the voltage is 10% below
the rated coil voltage the frequency should be no more than 15% above the rated coil frequency.

If the relay contacts and contactor coil, shown connected between the load side of the auxiliary contacts
and neutral, were connected directly between line and neutral from the generator the contacts would
chatter (repeatedly open and close) in the event of a short-circuit or severe overload. This is because
the voltage will fall and, when it is insufficient to energize the contactor coil, the contacts will open and
isolate the load. With no load connected the voltage will build up, until the contactor coil is reenergized, and the contacts will close into the severe overload and cause the process to repeat itself.
The repeated
opening and closing of the contacts will rapidly damage them. The circuit shown in Figure A7 prevents
repeated opening and closing of the contacts, once the start switch is in the run position, since if the
voltage falls sufficiently to de-energise the coil the auxiliary contacts will open and prevent reenergization of the contactor coil.
The switch, SI, is initially set to the start position in order to apply voltage to the contactor coil. Provided
that an overvoltage has not occurred, the d.c. relay contacts will be closed and the contactor coil will
become energized. The main and auxiliary contacts will close and when the switch is set to the run
position the contactor coil will remain energized. Contact chatter can occur when the switch is in the
start position.
This can be prevented by isolating the loads at start up with switches between the contactor and loads.
However, the load must then be introduced in steps as application of full load is likely to collapse the
voltage and cause the contacts to open. For three-phase systems, provided that no three-phase loads
are connected, each phase can be switched separately, with the voltage being returned to rated voltage
between switching operations by increasing the turbine power output.
For a single phase system, unless the transmission system can be conveniently split so that there is more
than one line, it will be necessary to switch full load directly with the contactor. In this case the turbine
must be started rapidly to minimise contactor wear and the contactor should be overrated by a factor of
at least two.
The d.c. relay and start switch should be rated for a current of at least 10 Amps, since the inrush current
into the contactor is quite large and highly inductive.
The voltage sensing circuit is the same for single-phase and three-phase systems. Voltage sensing
between just one line and neutral is adequate for a three-phase system, because even with
considerable phase inbalance the variation between phase voltages is quite small. The sensing circuit
has to be able to withstand the full overvoltage that occurs when the load is disconnected and the
turbine-generator overspeeds.
This should only last for a short time since the excitation capacitors will be switched out by the MCB(s),
as explained in Chapter 8. However, in case the MCB(s) fail to operate, the circuit has been designed to
withstand a continuous a.c. voltage of up to 600 V, which is higher than the maximum runaway voltage
for all but very large generators, as explained in Chapter 8.
KA1
KM1
Figure A 7 Circuit diagram for overvoltage trip.

The two power resistors are selected to set the operating voltage range for the trip. The 22 kQ resistor
will be used for all trips. The 2 Watt resistor can be selected with any value up to 4.7 kQ, the higher the
value the greater the voltage required for the trip to occur.
In series with the two power resistors is a variable resistor which allows some on site adjustment to be
made. If no on site adjustment is required then a shorting link can be used instead of this component.
The value of the variable resistor has been restricted to 2 kQ. for two reasons:
1) It prevents the trip voltage from being set too high, which could be damaging to the loads.
2) It prevents overheating of the variable resistor, since devices with a power rating above 0.5 W are
rare and expensive.
Due to the tolerances of the resistors and the operating voltage of the relay, the voltage range of the
trip should be determined by measurement. As a guide, with just the 22 kii power resistor in circuit the
tripping voltage range will be approximately 230 V to 250 V, depending on the setting of the variable
resistor. Higher voltages can be achieved by adding a 2 Watt resistor of suitable value.
The zener diodes are used to protect the variable resistor, capacitor and relay from being burnt out due
to an overvoltage. The resistor and capacitor across the relay contacts are to reduce arcing when the
contacts operate.
A relay with an a.c. coil would appear to be the obvious choice for this circuit. However, these exhibit
appreciable contact chatter when the voltage just reaches the operating voltage of the coil and
therefore they are not suitable. A d.c. relay has been used instead. A 110 V d.c. relay with a coil
resistance of 19,000 Q has been chosen, since lower voltage relays require a higher operating current
and therefore produce greater power dissipation in the power resistor and variable resistor.
Unfortunately relays with such a high coil voltage and resistance are difficult to obtain. However, a full
kit of components, including the circuit board, can be purchased through ITDG.

Appendix 6:
Power factor correction for motor starting
The unit described here is for single-phase motors. The circuit could be modified for three-phase motor
starting, though generally star-delta starting is more appropriate, as explained in Chapter 10.
The circuit shown in Figure A8 connects the power factor correction capacitor to the supply for just 0.5
to 1.5 seconds, depending upon the setting of the variable resistor. For most applications, this is
sufficient time to start the motor. If necessary the circuit values can be changed to increase the on time
of the capacitor. However, if the capacitor is connected for too long the motor will become
overcorrected for power factor and an overvoltage could be produced.
The capacitor should be sized to fully power factor correct the motor at the instant of starting. A large
value of capacitance is required. However, due to its intermittent use, a 'motor start' type capacitor can
be used. The circuit is connected after the supply switch, as shown in Figure A9.
The switching element which connects and disconnects the start capacitor is a d.c. operated solid-state
relay. These devices have a higher current rating for intermittent operation than for continuous
operation, as specified in the manufacturers' data sheets.
When the supply is connected the 33 ^F capacitor, CI, is rapidly charged via the diode bridge and
presents a d.c. source to the RC circuit to the right of it. The resistive part of this RC circuit is provided by
the fixed resistor R3, the variable resistor VR1 and the resistance in the input circuit of the solid-state
relay, as shown in Figure A10.
The solid-state relay (SSR) is treated as a current controlled device. When power is supplied to the
circuit, by switching on the motor, a current will flow through the input side of the SSR charging C2 and
causing the SSR to switch in the power factor correction capacitor. As C2 charges, the current flowing
gradually falls and the SSR will turn off, disconnecting the excitation capacitor.
With the component values shown, this takes approximately 0.5 seconds for VR1 equal to 0 Q and 1.5
seconds for VR1 equal to 47 kQ. However, much will depend upon the minimum operating voltage of
the SSR. Some indication of the switching time can be obtained by connecting an analogue ammeter in
series with the power factor correction capacitor. If a storage oscilloscope is available then the time
delay can be measured accurately.

Lo-SUPPLY IN No-POWER FACTOR CORRECTION UNIT TO MOTOR


Figure A9 Connection of power factor correction unit
1 kQ

Figure A10 Typical input circuit for a d. c. operated solid-state relay


The SSR should be rated for twice the peak supply voltage, since when the capacitor is switched out it
will initially remain charged and cause a peak voltage of twice the supply across the SSR output. In
addition, the peak input voltage to the solid-state relay must be less than its maximum rating. The
maximum peak input voltage is given by:
Vw_^Vn,-R8SR
(R3+RSSR)
'PK
Where V^ is the r.m.s. supply voltage
RSSR is the input resistance of the SSR
Note that when using this circuit with fridges and freezers and any other loads that contain a thermostat
the circuit must be connected after this switching element in order for it to operate every time the
motor is switched in.

The Polyphase Motor/Generator


OK, if you go through the whole thing, this will seem perhaps a bit long winded and maybe a bit spotty
description of how permanent magnet motor/generators work. But, the good news is if you just want to
know how to figure out the number of coils you need for a given number of magnets and how to hook
them up to successfully produce your own single-phase or polyphase motor/generator, you're in luck!
We'll cover that right off the bat with a set of simple instructions and formulas. What follows the basic
"how-to" information just fleshes out the details of what you're doing and why you're doing it.
One caveat for the rote instructions. They do assume you know, or know how to determine, the polarity
of your coil leads for proper connection. Store bought coils are likely marked for polarity. If you plan to
wind your own coils, (and why wouldn't you and miss all the fun?!), then here, without explanation, is

how to establish their polarity. Wind them all identically, all coils with turns in the same direction and
filling the form you wind them on in the same way. Choose the end of the wire you started with and,
arbitrarily, call it the positive (+) polarity lead, and call the other end of the wire coming from the coil the
negative (-) lead. Don't worry about which way you start making the turns in your coils, just do what is
easiest for your set up. Just be consistent from coil to coil, and all will be fine. Why consistency without
measurement works, we'll cover later.
Now, the instructions and formulas given below don't give all possibilities for coil connection and magnet
arrangement you might encounter. But, they do provide the information you need to make practical,
working, permanent magnet single-phase and polyphase motor/generators.
The same formulas work whether you're building a single-phase or a polyphase motor/generator. But,
they become so simple in the single-phase case that we'll cover the single-phase motor/generator
separately from the polyphase motor/generator.
Before getting to the actual construction instructions, first let's take a look at a diagramming method to
show motor/generator coil connections. The technique, presented in the section that follows, is adapted
from some of the figures in a well know book on electric motor repair by Robert Rosenberg [3]. A few
other books on the subject of permanent magnet motor/generator design and motor design in general are
listed in the references [4, 5,6].
Diagramming motor/generator phase coil connections:
Once you're decided on how many coils to use in your motor/generator, draw a small rectangle, one for
each coil, in a row across a piece of paper. To represent the coil leads draw a short line down from the
bottom corners of each of your rectangles. You can, but it isn't absolutely necessary, make a dot near the
left lead line to indicate the positive polarity coil lead for each rectangle. And, for each phase in your
motor generator, label the coil rectangles in a repeating sequence.
Traditionally, phases are labeled by the letters of the alphabet. So, given, for example, nine coils to use in
a 3-phase system you would repeat A,B,C in sequence three times, as A,B,C,A,B,C,A,B,C. Or, if you
were designing, say, a 5-phase system, you would label your boxes, in sequence, with the letters A
though E. If you like, you can identify each phase coil group by appending numbers to the letters. In our
case of nine coils in a 3-phase system, the coil box labels would then be, in sequence,
A1,B1,C1,A2,B2,C2,A3,B3,C3.
Now, regardless of whether you plan to build a single-phase or a polyphase motor/generator, due to the
alternating north and south poles of the rotor magnets, when operating, the direction of current through
each coil will be the opposite of the direction of the current through the coil that precedes it. (This is the
origin of the term Alternating Current (AC). More on that in the detail sections that follow.) So, under
each coil box, between the lead lines, draw a small arrow to represent current direction. Start from the left
box with an arrow pointing to the right, and alternate arrow direction with each box that follows.
And, that's it! The complete starting point diagram for a nine coil, 3-phase motor/generator is given in
figure 1. For a motor/generator with more or less coils, just draw more or less boxes.
Armed with the base of a diagramming technique for producing a simple picture of what we want to
build, we can now get on with how to actually find the number of coils we need, and how to hook them
up.

Coil count and connections for a single-phase motor/generator


1. Dig through your junk box and find all your magnets.

2. Make piles of the same types of magnets


3. Choose the pile containing the type of magnet you want to use in your single-phase motor/generator.
4. Count the magnets in the chosen pile. If it is an odd number, toss one magnet back in the junk box, or go
find another one, so that you end up with an even number of magnets.
5. Make the same number of coils as you have chosen magnets.
6. Lay your coils out in a row so that their positive lead is on the same side of each coil in the row, and
connect them in series so that the negative lead of the first coil connects to the negative lead of the
second coil, the positive lead of the second coil connects to the positive lead of the third coil, and so on.
The proper connection is simple to show using the diagramming technique presented above. Just draw
lines to connect the coil boxes so that the lead nearest the head of one box's arrow connects to the lead
nearest the tail of the next box's arrow (figure 2). Physically making this connection relative to the
positive leads as drawn on your coil diagram guarantees the alternating current flow directions in the
coils do not oppose each other.

7. Build your motor/generator with magnets alternating north and south poles evenly spaced around it's
rotor, and coils evenly spaced around it's armature, as, for example, shown in figure 3 for an eight
magnet, 8 coil, single-phase motor/generator.

OK, there is a bit of a misleading implication above. In steps (6) and (7) above there is no differentiation
between connecting coils for a single-phase motor or for a single-phase generator. The connection will
work just fine as described for a generator. And, in fact, as described, the connection will also work for a
motor. Connect an AC power source to a motor/generator with an armature wired as in figure 2, and it
can be made to spin via the input power. But, it will not self start. You will need to give it a bump to get it
to go. A single-phase motor can be connected to self start. We'll leave that discussion until after we talk
in more detail about polyphase motor/generators.
Coil count and connections for a polyphase motor/generator:
Dig through your junk boxes and find all your magnets.
Make piles of the same types of magnets
Choose the pile containing the type of magnet you want to use in your polyphase motor/generator.
Count the magnets in the chosen pile. If it is an odd number, toss one magnet back in the junk box, or, go
find another one, so that you end up with an even number of magnets.
Decide how many phases you want to use in your polyphase motor/generator. Most everyone on the
planet chooses three, and, I'm going to choose 3 for this discussion, so you might as well, too. You don't
have to choose three phases, but at least choose an odd number. (To show this technique does generalize
I'll diagram hooking up a five-phase motor/generator after the 3-phase discussion.)
Calculate the number of coils required:
For a given number of phases, N, you need N coils per magnet pair in your N-phase motor/generator.
For a count of M magnets (often referred to as poles), with M an even number, we have M/2 magnet
pairs.
So, for an N-phase motor/generator with M poles the number of coils required, C, is:
C = N*(M/2).

I found 8 magnets in my junk box when I set out to write this. So, for a 3-phase motor/generator the
number of coils we want to wind is:
3*8/2 = 12.

Calculate the number of coils per phase:


For C coils and N phases, the number of coils per phase is C/N. Here, for our 8 pole, 3-phase
motor/generator we get 12/3 = 4 coils per phase. This means you will be connecting three groups of 4
coils each in your motor/generator.

Select the motor/generator wiring configuration:


For a 3-phase system, there are two primary types of polyphase motor/generator connections, Delta, and
Y (also know as a Star connection). For a Delta connection, the phases are connected in a triangular
configuration. In a Y connection the phases are connected in 3-armed, or "Y" shaped configuration. The
Delta configuration is more efficient for low rpm operation, and we'll choose that for now. The Y
configuration will be covered in the details discussion that follows these rote instructions and formulas.
Diagram the phase connections:
Using the previously described diagramming technique (figure 1), each of the four coils in the 3 phase
groups is connected in series just as in the single-phase motor/generator so there is proper current flow
through the group. Draw lines so they connect (+) to (+) and (-) to (-) polarity coil connections for each of
the phase coil groups (figure 4). You can use the "follow the arrow" method with each phase group the
same as described in step (6) for the series connection of all coils in the single-phase motor/generator.
Note carefully how the B phase starts on phase coil B2, not coil B1. This is because the first B phase coil
who's current direction matches the current direction of the first coils in the A and C phase coil groups is
B2, not B1.

Once the A,B, and C, phase coil groups are connected, then the groups are connected to form the chosen
Delta or Y configuration. For a three-phase Delta connection the end lead of the A phase coil group is
connect to the starting lead of the C phase coil group, the end lead of the C phase coil group is connected
to the start lead of the B phase coil group, and the end lead of the B phase coil group is connected to the
start lead of the A phase coil group, as diagrammed in figure 5. Again, as pointed out in step (9) above,
note the starting lead for the B phase coil group comes from coil B2, not coil B1.

Build it. Figure 6 shows an example physical diagram for a three-phase, 8-pole, 12-coil, Delta connected
motor/generator
.

If you use your motor/generator in motor mode and want it to spin in the opposite direction from what
you will find with the motor/generator design given here, you can reverse any two of the input phase

connections, and the motor will run in reverse. We'll see why that is later. (And, of course, you could also
physically turn the motor around 180 degrees so that the end of the shaft facing you becomes the end of
the shaft pointing away from you.)
There is no issue with self starting of polyphase motors. Connect three-phase power to a 3-phase
motor/generator and it will run as a motor without any of the startup connection tricks required for a self
starting single-phase motor to be discussed later.

Of magnets and wires and such:


Getting into the spin:
First thing, magnets. Most everyone has played with magnets at one time or another. Point the opposite
poles of two magnets at each other, north to south, and they attract. Point the same poles at each other
(north to north, or south to south) and they repel. Interaction of the magnetic fields produced by the
magnets is what causes the attraction and repulsion effects observed for a pair of magnets.
The same as it is for the magnetic field of the earth, the magnetic field of a simple bar magnet is
described as lines of magnetic flux which extend from a magnet's north pole to its south pole (figure 7).
(It might be good to note here that the north geographic pole of the earth is actually a south magnetic
pole. Which explains why the north pole of the magnet in a compass points toward the earth's geographic
north pole.)

Considering the magnet flux lines as drawn in figure 7, magnetic attraction and repulsion can be viewed
as the interaction of the arrows (or vectors) giving the direction of the flux lines from north pole to south
pole. When two north or two south poles are brought next to each other, the arrows from one magnet's
pole point in the opposite direction of those from the other magnet's pole, and, hence, "collide," pushing
each other away, causing the repulsion effect. But, when a north pole and a south pole are brought
together, the arrows point in the same direction, and one set "sucks" the other set along, much like two
streams of water flowing in the same direction, which causes the attraction effect. (This is, of course, just
a description for visualization purposes.)
Magnetic repulsion and attraction are the basis of operation for an electric motor. By proper alignment
and timing of magnetic fields, the parts of an electric motor can be made to push and pull on each other
so that a smooth continuous motion is obtained.
Besides having a magnet handy, another way to generate a magnetic field is to run an electric current
through a piece of wire. The flow of current through a wire produces lines of magnetic flux around the
wire just like the lines of magnetic flux produced by a magnet. The lines of flux are produced such that
when you point your left thumb along the wire in the direction of current flow through the wire, your left
fingers will curl around the wire in the same direction as the lines of flux around the wire, i.e., you finger
tips will point in the same direction as the arrows seen in a diagram of the flux lines (figure 8).

Obviously, if the direction of current through the wire is reversed, then you have to point your left thumb
in the opposite direction, and, as shown by the curve of your fingers, the lines of flux around the wire will
be in the opposite direction, effectively changing the polarity of the magnetic field around the wire.
We aren't going to get deeply into what constitutes the "real" direction of current flow in a wire. As has
been pointed out above, consistency is what counts. Make your favourite assumption, use it without fail,
and all will be well.
Here we'll consistently consider current flow to be from the negative terminal of a power supply to the
positive terminal of a power supply. That means we're looking at electron flow, and not current flow as it
was originally assigned to be from positive to negative for the first electric batteries, (a concept based on
Benjamin Franklin's earlier description of two types of electricity, "positive" and "negative," which, well,
frankly, some would say he got backwards). Engineers and physicists can argue about what is really
going on all they want, and we'll just get on with the task at hand. Symmetry, which we'll talk about more
a bit later, is the key to why, for practical purposes, the initial choice doesn't really matter. In fact, if you
are more comfortable considering current flow to be from positive to negative then consistently use that
idea, and instead of using the left-hand rule as described above, just use your right hand to apply the
right-hand rule. Then following the curve to your right finger tips will get you the same results as we've
seen for the left-hand rule. That's symmetry!
Now, the nature of magnetic poles around a straight piece of wire with current flowing through it isn't
immediately obvious. But, if you turn the wire into a loop, the orientation of magnetic poles relative to
the wire becomes more clear. As shown in figure 9, with current flow from left to right and the loop
directed into the page, the flux lines circulate into the top of the loop and out of the bottom. Comparing
this to the flux lines seen for the magnet in figure 7, we can see this implies the north pole of the loop's
magnetic field is at the bottom, and its south pole is at the top.

The effect becomes more clear when there are multiple loops stacked on top of each other to form a coil.
Since the flux lines produced by each loop combine in the same direction they produce a larger (and
stronger) magnetic field (figure 10). This is why more turns in the coil of an electromagnet make for a
stronger electromagnet.

As can be seen in figure 10, the same repulsion and attraction of flux lines can be had from the coil as
from a real magnet. By reversing the direction of current through the coil the orientation of its north and
south poles is reversed. And, by taking this action in to account, simple electric motors can be devised.
For example, with proper timing of the change in direction of current through two coils, a magnet on an
axle can be made to rotate between the coils (figure 11).

Going the other way:


OK, in the last section we made it as far as a good idea of how to make a simple electric motor. A lot of
details need to be filled in yet. Like, in particular, how to control switching and timing of the magnetic
fields, but it's a definite start. And, with the information we put together there about magnets and wires
and such, we are in good shape to move ahead with an initial description of how electric generators work.
As mentioned before, the aspects of physics we're looking at are symmetric. So, just as running a current
through a wire creates a magnetic field, moving a magnet near a wire creates an electric current in the
wire (figure 12).

Of course, the more wires a moving magnetic field crosses, the more currents that are generated, one for
each wire (figure 13). And, when a wire is wound in a coil, the effect is the same for each loop in the coil,
except, because the coils are connected, the multiple currents created as the magnetic field passes over the

coil windings add together. So, the more windings in the coil, the more current generated as the magnetic
field passes by (figure 14). This is the basic idea behind an electric generator.

Also, the stronger the moving magnetic field, the more current generated in any wire it moves by. So,
either by putting more turns in its coils, or by using stronger magnetic fields, or both, the more power that
can be produced from a generator or utilized from a motor.
We've already see examples of making simple permanent magnet generators in the coil count and
connection sections above. So, let's continue on and investigate a few more details that will lead us to an
understanding of why we make the connections we do as previously described for single-phase and threephase delta motor/generators, and also how to define connections for other than single and three-phase
systems.
Pick a phase, any phase:
OK, so what does "phase" mean, anyway? Well, it's another one of those words that changes meaning
with context. Regarding polyphase motor/generators, it is often used interchangeably to describe two

main features. It can mean one of the multiple leads (or the lead's associated electrical waveform) in a
polyphase system, or it can mean the difference (measured in degrees) between the peaks seen in the
waveforms found in any two leads in a polyphase system. For the second case that difference is more
correctly referred to as the "phase angle," which, here, is a term that relates to the notion of a rotating
magnetic field. The rotating magnetic field concept is at the heart of any motor/generator system, be it
single or polyphase.
That we are dealing with rotating magnetic fields seems easy enough to accept. The devices we've been
discussing have magnets that spin on an axle. So, there it is, rotating magnetic fields. Pretty unavoidable,
that. Of course, as always, the devil is in the details.
Sinusoids:
Rotation implies circles. But, to draw a circle covering the 360 degrees of each rotation to represent the
motion over time of a motor/generator that may be spinning at thousands of revolutions per minute
wouldn't be too informative. So, instead, we draw a sinusoid, which is just a way of displaying circular
motion in a linear diagram.
To see this, consider a simple machine consisting of a disc with a peg at its edge parallel the disc's center
axle, and a "T" shaped piece which has a slot in its crossbar. The T-bar is laid on the disc so that the
crossbar slot fits over the peg, and the upright points down when the machine is looked at in plan view. A
guide aligned with the center of the disc is placed over the T-bar upright so that while the peg on the disc
can slide back and forth in the slot in the T-bar crosspiece as the disc rotates, the T-bar upright can move
up and down, but not side to side. So, while the disc rotates, the peg moves in a circular motion which
forces the T-bar to move so that its end travels up and down a distance equal to the diameter of the circle
defined by the motion of the peg (figure 15).

If a pencil is attached to the end of the T-bar upright piece so that it will leave a mark on a piece of paper
placed under it, then, as the disc rotates the pencil will draw a straight line with length equal to the
diameter of the peg circle each time the disc rotates. We can spin the disc until the pencil wears out, but
that is all we will see, a single vertical line. However, if we move the paper parallel to the T-bar cross
piece, we'll see something quite different, that is, a repeating curve, called a sinusoid, which gives the
diameter of the peg circle via its amplitude, and repeats its waveform across the paper through a number
of cycles that relates to the rotational speed of the disc and the speed of motion of the paper under the
pencil
If we know the speed of the paper, then we can deduce the rotation speed of the disc by counting the
number of cycles in a given distance across the page, equating distance to the time it took the paper to
travel that distance, and dividing the number of cycles see in the sinusoid over the given distance by the
equivalent time for that distance. The result is disc rotation speed expressed as frequency in units of

cycles (rotations) per chosen time unit.


If we arbitrarily designated the start (0 degrees) position as when the peg is at its topmost position when
the disc rotates, and start moving the paper to the right just when the peg reaches that point, then the
pencil on our simple machine will trace out a sinusoid that represents the circular motion of the disc as a
"wave" that has a positive maximum value at the start time position, drops to zero when the disc has
rotated 90 degrees, reaches a maximum negative value as disc rotates through 180 degrees, returns to 0
when the rotation reaches 270 degrees, and climbs back to the maximum positive value as the rotation
goes though 360 degrees. The 360 degree rotation returns the pencil to the maximum positive position,
but a distance across the page that equates to one rotation time, and the cycle repeats (figure 16).

The time for a sinusoid to complete one rotation cycle is referred to as its period (T). The number of
cycles the sinusoid completes in a fixed time period is referred to as its frequency (f). Traditionally the
time period to count cycles to determine frequency is one second, and frequency is expressed in cycles
per second or Hertz (Hz). Note that frequency is the inverse of period, that is, divide 1 by the period
value, and you get the frequency value, or, divide 1 by the frequency value and you get the period value:
.

Rather than using spinning discs with sliding pencils to diagram rotations, the trigonometric function
cosine has been defined to represent just this motion. In a mathematical formula the cosine function is
given as cos(x), where the argument x represents the degrees of rotation. The cosine function follows the
same form of curve as described by the pencil and disc arrangement of figure 16; having a value of 1 for
an argument of 0 degrees, a value of 0 for an argument of 90 degrees, a value of -1 for an argument of
180 degrees, a value of zero for an argument of 270 degrees, and returning back to 1 for an argument of
360 degrees.
The amplitude of the curve drawn by the pencil and disc method is equal to the diameter of the disc. But,
since we have defined things to range over plus and minus values, and the cosine function ranges from 1
to -1, we multiply the cosine function by half the diameter (the radius) of the disc to get the proper values
for the curve relative to the degrees argument to the cosine function. That is, if the diameter of the disc is
A, then at any particular degrees of rotation value, x, the proper height, h, of the sinusoid is given by:

We can't make the mistake of thinking that the diameter of the disc mentioned above has anything to do
with the diameter of our motor/generator rotor disc. The "diameter" we are really talking about is simply
the amplitude of the sinusoidal electrical waveforms applied to, or generated by, our motor/generator. In
fact, without any loss of usefulness, we can drop the division by 2 in the amplitude factor, and just use the
parameter A alone to indicate we want to allow sinusoids with an amplitude range other than from -1 to
+1, assume the value of A will be chosen correctly when necessary, and go with the slightly simpler
equation:
Now we almost have a mathematical formula that can be used to diagram the phase waveforms seen in
our motor/generators. But, there is still that matter of x, the unknown angle. If we just wanted to plot one
sinusoid, then varying x from 0 to 360 and plotting the result for h as the y value in an x-y plot would do
fine. But, that isn't going to give us what we need when it comes to looking at the phase angle between
phase waveforms. Also, often we won't want to plot with degrees on the x axis of our plots, but rather
with time. Fortunately, since we are dealing with known rotation speeds, (i.e., the frequency of the
sinusoidal waveforms in question, e.g., the 60 cycles per second (60 Hz) for commercial power found
around here), that's relatively easy to figure out.
So, back to considering imaginary spinning discs and sliding pencils for a bit. Another way of measuring
angles is in radians. There are
radians in 360 degrees (making one radian approximately 57.297
degrees). So, one rotation of the disc in figure 15 means the peg which moves the crossbar sweeps
through an angle of
radians. Radian measure is just a scaling factor different from degree measure,
and the cosine function can be defined to use radians as an argument as well as degrees. For the time
being we'll use that definition. It might seem an added confusion to switch from degrees to radians, but,
that is really just a matter of what you are used to. We gain something by switching to radian measure,
that is, when coupled with the concept of radial velocity, the ability to use time as the argument to the
cosine function.
Radial velocity ( ) is a measure of the speed of rotation of a disc in terms of radians per second, rather
than cycles per second. Again, we aren't really talking about speed of rotation of a disc, but the "speed"
with which a sinusoidal waveform completes its repeated identical excursions. The concept of radial
velocity still applies, since that sinusoid has the same circular motion related form as the one produced by
the disc and pen method of figure 15. In either case, multiplying the radial velocity by time (measured in
seconds) gives us radians as a proper argument to the cosine function, and, thus, knowing the radial
velocity, we can plot our sinusoid against time on the x-axis of an x-y plot, rather than degrees. That
brings us to a new form for our equation:
A bit more work with this expression will lead us to the ability to check the phase angle between
sinusoids at a particular time, which is a lot easier to work with than trying to check phase angles related
to portions of a cycle, which would be the case if we used degrees and frequency in our equations, rather
than radians and radial velocity.
Before we move on to checking phase differences, there is one more important relationship to note, the
connection between frequency and radial velocity. The unit Hz (cycles per second) refers to how many
rotations per second we see for our rotating disc or how many full excursions per second we see in our
sinusoidal wave form. In either case, one rotation or one sinusoidal excursion, we can equate the
frequency to sweeping an angle of 360 degrees with every cycle. So, for example, we could say that a 60
Hz signal is also a 60*360 = 21600 degrees per second signal. Of course, that isn't a very useful piece of
information. We need some way to relate the degrees of sweep per second for a signal frequency
measured in cycles per second to radial velocity so that we can apply the value we know, frequency, in
our equation that uses radial velocity and time. Radians to the rescue!
To use frequency in our equation in and t, we need to convert from frequency to radial velocity so that
the multiplication by t will give the proper value in radians. That we can do by noting, as pointed out
before, one revolution is
radians, and also that one cycle in frequency is one revolution, then multiply
the appropriate factors together so that the units cancel out to give radians per second, the proper units for
, i.e,

That leads us to another form for our sinusoidal equation:


,

which allows us to plot our waveforms against time, while using frequency in the argument to the cosine
function.
So, all that verbage just to explain a factor of
? Well, yes and no. The warm up to expressing circular
motion as a sinusoid will prove useful later, when we look at making connections in a polyphase
motor/generator in terms of vector sums. At least we are done here now, and can get on with the next
phase in our discussion.
Phase vs. phase:
I'll state in advance the discussion here may get a little confusing. This is because, as mentioned before,
phase is one of those words that changes meaning with context. The main uses for "phase" here will be to
mean one of the input/output wires to a motor/generator or one of the phase coil groups as described
earlier in the polyphase coil count and connection section (these two meanings are basically
synonymous), or also to mean the sinusoidal waveform found in an input/output wire or phase coil group.
Another potential confusion may come in use of the terms "phase difference" and "phase angle." The
phase difference is the difference between phases. (Well, duh!). That difference is usually measured as an
angle, and, so, it is often referred to as phase angle. More correctly, the phase angle is the value of the
phase difference, but, whatever, the two terms are used interchangeably elsewhere, and will be here, too.
So, OK, that being that, we're just about there. We have a cosine equation we can use for generating
sinusoids, in which we can use variables we know or can determine, time and frequency. Now we need a
way to include phase difference into our equation.
The phase difference between two sinusoids is the separation in angle between the same relative point on
the waveforms. We measure the difference as an angle because the sinusoids are representations of
circular motion and the difference between two points on a circle is normally expressed as an angle. That
angle is not quite enough information to describe the difference. We also need to know which waveform
comes first. That association is described using the terms "lead" and "lag." When one waveform leads
another waveform then the first one reaches the end of a cycle before the second one. When one
waveform form lags another waveform, then the first waveform reaches the end of a cycle after the
second one (figure 17).

To include a lead or a lag in our cosine equation, we simply subtract or add a phase angle in the main
angle argument. Careful here. The relationship isn't quite intuitive. If we have two sinusoids rotating with
the same frequency, and only apply the phase correction to one of the two, then to force a lead by the
uncorrected waveform over the uncorrected waveform we subtract the phase angle from the corrected
waveform, and to force a lag of the uncorrected waveform we add the phase angle to the corrected
waveform (figure 18).

We've now reached the final form of our cosine equation:


,

where the phase angle, , may be positive or negative, and, as used here, should be expressed in radians.
If is provided in degrees, then multiply by and divide by 180 to convert degrees to radians, which
leaves us with the same basic form of the equation:
,

In a later section we will be introducing another multiplication factor for the argument in f and t. But, for
now we'll ignore that and move on to look at expressing the full set of phase angle relationships for a
polyphase motor/generator via our latest equation.
In general, to make a set of plots for phase comparisons, we need an equation for each sinusoid, with each
equation expressing the amplitude, frequency, and phase difference for its associated sinusoid. For
arbitrary sinusoids each equation would have a completely different set of arguments and multiplication
factors. For example, the equation set for three arbitrary sinusoids, A,B,C, would be, with the phase angle
expressed in radians:

Using these equations we could compare three sinusoids of arbitrary magnitude, arbitrary frequency, and
arbitrary phase angle at any point in time. But, here we make two simplifying assumptions, that the
magnitude of each sinusoid is the same, and the frequency of each sinusoid is the same. Note that with the
identical magnitudes assumption we can set the magnitude factor to 1 in each equation with no loss in
generality. Also, note that with the identical frequency assumption time is no longer a critical factor if we

are only comparing phase differences, and we can make the comparisons at any arbitrary rotation angle,
say, . We can make one more simplification that has its origin in standard practice rather than physics,
which is when comparing phases, we assume the phase angle for the first phase is zero, and adjust the
other phase angles accordingly. That gives us, in our three phase case, the following simplified equation
set:

So, now we just need the phase angles for the B and C sinusoids.
In a polyphase motor/generator the sinusoidal waveforms found in each phase have a fixed angular
relationship, determined by the number of phases. Each phase leads the next one by the same angle. That
phase difference, P, is simply one full cycle expressed as an angle, 360 degrees or
radians, divided by
the number of phases. Using degree measure:
P = 360/N.
Thus, for a three-phase system, we get a fixed phase angle of P = 120 degrees. Of course, it would be just
a little to simple if we could use 120 degrees, (or 120 /180 radians), for the phase angles in our
simplified equation set. If we draw three sinusoidal waveforms, A,B,C, each leading the next by 120
degrees, (figure 19) we can find the required phase angle values for our set of cosine equations by
examining the plots.

We know we want a phase angle of zero for waveform A. And, traditionally, we also use positive value
phase angles. That means we will look for appropriate phase lags to define our phase angles. The
positions where the phase A and B waveforms reach their maximum value within one full cycle, with A
lagging B, is at = 120 degrees for B and = 360 degrees for A. That gives us a phase difference
between A and B of 360 - 120 = 240 degrees, which we use as the phase angle for the B waveform cosine
equation. Similarly, the phase difference for A lagging C is 360 - 240 = 120 degrees, which we use as the
phase angle for the C waveform cosine equation. That gives us the following equation set:

which is the standard form for three-phase equations. To determine lag values from the given phase lead
differences of 120 degrees without having to plot the waveforms, we can just note, as pointed out earlier,
(figure 18), that a phase lead relationship between two waveforms can be reversed by subtracting the
phase difference from 360 degrees. So, for A and B, with A leading B by 120 degrees we can also say A
lags B by 360 - 120 = 240 degrees. And, since there are 360 degrees in one cycle, if A lags B by 240
degrees and B leads C by 120 degrees, then A must lag C by 120 degrees. Similar analyses can be made
for systems with other than three phases.

From a strictly theoretical point of view, that's it! We have now have our complete set of cosine equations
for three-phase related sinusoids, which we can use in terms of angle of rotation as in the three equations
above, or in terms of time via the radial velocity, which, as we have seen, can be expressed directly, ,
or in terms of frequency, f, if we include the proper conversion factor of
. If we also include an
amplitude factor, K, and convert the degree value phase angles to radian values by multiplying by and
dividing by 180, we get a general set of equations we can use to compare three-phase waveforms of any
magnitude and any frequency at any time:

We'll make use of these equations in one from or another later, when we look at why we make the
connections we do between phase coil groups in a polyphase motor/generator.
Mechanical vs. electrical degrees:
A while back I mentioned we'd look at one more factor to apply to the main argument in our cosine
equations. And, now the time has come. Here we'll step back a bit from the predominantly theoretical
discussions and look at a result that stems from the basic physical arrangement of magnets and coils in
our motor/generators. This factor is needed to properly define the phase sinusoids from our cosine
equations if we use frequency or radial velocity to mean the rotational speed of our motor/generator rotor.
If we use frequency or radial velocity to mean the actual cycle rate of our phase sinusoids, then this factor
is not needed in the cosine equations. This discussion assumes we are relating our cosine equations to
rotor revolutions. When we relate our equations to the actual phase cycle rate, (e.g., 60 Hz for US
commercial electrical power), then the last set of three equations above are the ones we want to use for a
three-phase motor/generator.
While it is true we relate cycles in sinusoids to circular motion, and here, spinning of the rotor in a
motor/generator does ultimately define that relationship, one rotation of our motor/generator rotor does
not equate to one rotation in our phase sinusoids. In fact, for an AC generator, each turn of the rotor will
result in a number of sinusoidal cycles in the output phase waveforms. Similarly, a number of input
power cycles are required to produce a single turn of the rotor in an AC motor/generator used in motor
mode.
Physical turns of a rotor, as well as cycles of a sinusoidal signal can both be measured as an angle in
degrees or radians. For the moment we'll use degrees. To differentiate the two motions, we call the
angular measure of turns of the rotor "mechanical degrees," and the angular measure of cycles of the
sinusoidal waveforms "electrical degrees." For a particular motor/generator design, there is always a fixed
ratio between the two.
For the model motor/generator described in the first part of this web page there are eight sinusoidal phase
cycles for each turn of the rotor. Links to the source code, (you might want to check the comments), and
it's outputs, for a program that computes and plots the electrical degree measure phase waveforms for this
motor/generator are provided below:

Output data file: RAM:polephase.data


pole rotation step: 1.500 degrees

step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step

0:
1:
2:
3:
4:
5:
6:
7:
8:
9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25:

deg-m

deg-e

0.000
1.500
3.000
4.500
6.000
7.500
9.000
10.500
12.000
13.500
15.000
16.500
18.000
19.500
21.000
22.500
24.000
25.500
27.000
28.500
30.000
31.500
33.000
34.500
36.000
37.500

0.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000

Phase A

Phase B

Phase C

1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500

-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000

-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500

step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step

26:
27:
28:
29:
30:
31:
32:
33:
34:
35:
36:
37:
38:
39:
40:
41:
42:
43:
44:
45:
46:
47:
48:
49:
50:
51:
52:
53:
54:
55:
56:
57:
58:
59:
60:
61:
62:
63:
64:
65:
66:
67:
68:
69:
70:
71:
72:
73:
74:
75:
76:
77:
78:
79:
80:
81:
82:
83:
84:
85:
86:
87:
88:
89:
90:
91:

39.000
40.500
42.000
43.500
45.000
46.500
48.000
49.500
51.000
52.500
54.000
55.500
57.000
58.500
60.000
61.500
63.000
64.500
66.000
67.500
69.000
70.500
72.000
73.500
75.000
76.500
78.000
79.500
81.000
82.500
84.000
85.500
87.000
88.500
90.000
91.500
93.000
94.500
96.000
97.500
99.000
100.500
102.000
103.500
105.000
106.500
108.000
109.500
111.000
112.500
114.000
115.500
117.000
118.500
120.000
121.500
123.000
124.500
126.000
127.500
129.000
130.500
132.000
133.500
135.000
136.500

312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000

0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978

-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309

0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669

step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step

92:
93:
94:
95:
96:
97:
98:
99:
100:
101:
102:
103:
104:
105:
106:
107:
108:
109:
110:
111:
112:
113:
114:
115:
116:
117:
118:
119:
120:
121:
122:
123:
124:
125:
126:
127:
128:
129:
130:
131:
132:
133:
134:
135:
136:
137:
138:
139:
140:
141:
142:
143:
144:
145:
146:
147:
148:
149:
150:
151:
152:
153:
154:
155:
156:
157:

138.000
139.500
141.000
142.500
144.000
145.500
147.000
148.500
150.000
151.500
153.000
154.500
156.000
157.500
159.000
160.500
162.000
163.500
165.000
166.500
168.000
169.500
171.000
172.500
174.000
175.500
177.000
178.500
180.000
181.500
183.000
184.500
186.000
187.500
189.000
190.500
192.000
193.500
195.000
196.500
198.000
199.500
201.000
202.500
204.000
205.500
207.000
208.500
210.000
211.500
213.000
214.500
216.000
217.500
219.000
220.500
222.000
223.500
225.000
226.500
228.000
229.500
231.000
232.500
234.000
235.500

24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000

0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105

-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809

-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914

step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step

158:
159:
160:
161:
162:
163:
164:
165:
166:
167:
168:
169:
170:
171:
172:
173:
174:
175:
176:
177:
178:
179:
180:
181:
182:
183:
184:
185:
186:
187:
188:
189:
190:
191:
192:
193:
194:
195:
196:
197:
198:
199:
200:
201:
202:
203:
204:
205:
206:
207:
208:
209:
210:
211:
212:
213:
214:
215:
216:
217:
218:
219:
220:
221:
222:
223:

237.000
238.500
240.000
241.500
243.000
244.500
246.000
247.500
249.000
250.500
252.000
253.500
255.000
256.500
258.000
259.500
261.000
262.500
264.000
265.500
267.000
268.500
270.000
271.500
273.000
274.500
276.000
277.500
279.000
280.500
282.000
283.500
285.000
286.500
288.000
289.500
291.000
292.500
294.000
295.500
297.000
298.500
300.000
301.500
303.000
304.500
306.000
307.500
309.000
310.500
312.000
313.500
315.000
316.500
318.000
319.500
321.000
322.500
324.000
325.500
327.000
328.500
330.000
331.500
333.000
334.500

96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000
168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000
12.000
24.000
36.000
48.000
60.000
72.000
84.000
96.000
108.000
120.000
132.000
144.000
156.000

-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914

0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809

-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105

step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step
step

224:
225:
226:
227:
228:
229:
230:
231:
232:
233:
234:
235:
236:
237:
238:
239:
240:

336.000
337.500
339.000
340.500
342.000
343.500
345.000
346.500
348.000
349.500
351.000
352.500
354.000
355.500
357.000
358.500
360.000

168.000
180.000
192.000
204.000
216.000
228.000
240.000
252.000
264.000
276.000
288.000
300.000
312.000
324.000
336.000
348.000
360.000

-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500
-0.309
-0.105
0.105
0.309
0.500
0.669
0.809
0.914
0.978
1.000

0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500
-0.669
-0.809
-0.914
-0.978
-1.000
-0.978
-0.914
-0.809
-0.669
-0.500

0.309
0.500
0.669
0.809
0.914
0.978
1.000
0.978
0.914
0.809
0.669
0.500
0.309
0.105
-0.105
-0.309
-0.500

Those of you checking my work may have already noticed that the coil count for the motor/generator I
constructed doesn't match the number you would calculate using the rote formulas presented at the start
of this "How's it work?" section. That difference is due to consideration of a matter known as cogging,
where the drag of magnet fields across the coils affects motor/generator performance. Don't worry about
that, we'll get to it in a while. For now let's continue the discussion by looking at the relationship between
electrical and mechanical degrees for the eight magnet, 12 coil set we computed earlier for a 3-phase
motor/generator using the rote formulas.
Once again symmetry comes into the picture. Referring back to figure 6, we can define a mechanical 0
degrees position for the illustrated motor/generator as the middle of the topmost armature coil, C4. Now
say we turn the rotor so that a north pole magnet aligns with the mechanical 0 position, and define an
electrical 0 position to be where the north pole magnet is aligned with the mechanical 0 position.
Hmmmm...looks like if we turned the rotor to line up another north pole magnet with the mechanical 0
position, we'd have another electrical 0 position. That's right! And, that is the answer. The rotor is radially
symmetric, so when we rotate it, we can align it at several positions and see an identical picture.
In the previous paragraph we defined "a" mechanical 0 degree position and also "an" electrical 0 degree
position. Those nonunique statements, as opposed to defining "the" mechanical or electrical 0 degree
positions, were intentional. Because of how figure 6 was drawn it is easy to see coil C4 as defining a
mechanical 0 degree position. And, in fact, because of symmetry, the relationship we are about to
investigate could be arrived at using coil C4 to mark the mechanical 0 degree position. However, because
we want to look at what happens over a full sinusoidal phase cycle when defining the electrical 0 degree
position, we will start at the the beginning of a phase coil group, not because it is at all necessary, but just
because it looks better in the figures. We can start on any phase coil group. Let's choose group A, and
define the mechanical 0 degree position as being marked by coil A1, which, in figure 6, is just to the right
of coil C4.
Again we define the first electrical 0 degree position to be where a north pole magnet aligns with the
mechanical 0 degree position. This is the position where, when the motor/generator is in operation, the
phase A sinusoid will be at its maximum value. Now, as the rotor turns, there will be multiple positions
where different north pole magnets will pass under coil A1, and all other coil and magnet alignments will
be such that the phase A sinusoid will reach a maximum. Each time this happens phase A has passed
through a full sinusoidal cycle, and each of these alignments corresponds to a new electrical 0 degree
position. The number of full phase cycles per single mechanical rotation is always in a fixed ratio. That
ratio will vary with motor/generator design, but, we can always see what the ratio is by looking at the
relative alignment, in degrees, between coils and magnets in the design.
We can depict coil/magnet alignment in a linear diagram, similar to how earlier we diagrammed phase
coil connections as linear groups, even though a real motor/generator armature is circular (e.g., figure 4).
In our eight magnet, 12 coil system, there will be 360/8 = 45 degrees between magnets, and 360/12 = 30
degrees between coils. Drawing this arrangement linearly, with one north pole aligned with coil A1
(figure 20) we can see that the rotor will have to shift (rotate) 90 degrees to achieve the identical
alignment of magnet poles and coils as seen for the first electrical 0 degrees position.

If we see a new electrical 0 degree position for every 90 degrees the rotor turns then for every turn of the
rotor we will see 360/90 = 4 sinusoidal phase cycles. If you make the same analysis for a ten magnet rotor
and use the rote calculation number of coils, 15, for a 3-phase motor/generator, you will come up with a
ratio of 5 phase cycles per rotor turn. With a little squinting at the diagrams, the general relationship for
the ratio, R, of phase cycles to rotor turns can be seen to be the result for 360 degrees per phase cycle
being divided by one-half the number of magnets divided into 360 degrees per rotor turn. That is, if M is
the number of magnets:
R = 360/(360/(M/2)).

With a little algebraic manipulation, that becomes:


R = M/2,

which is just the number of pole pairs in the motor/generator. In fact, so long as the motor is properly
designed, the number of coils doesn't enter into the relationship. For example, we know the coil count for
the 16 pole motor/generator constructed for this web page is not what you would calculate from the rote
formulas. But, we also know the electrical to mechanical degrees ratio for that motor/generator is 8, and,
by our new formula:
R = M/2 = 16/2 = 8.

So, when we use rotation speed of our motor/generator rotor in the main angle term of our cosine
equations, we apply the ratio R = M/2 to that term to produce the proper number of phase cycles per rotor
revolution, giving us another form for our equations:

OK, I picked a phase, now what?


So, now we've got some phases spinning around, and described then in fairly painful detail. What do we
do with them? Why, we hook them up to make motors and generators, of course!
For a single-phase motor/generator, we've already seen that isn't too difficult. Well, except for that bit
about self-starting of single-phase motors. (Which we'll get to real soon now.) But, it gets a little more
tricky with polyphase motor/generators. There we have multiple electrical signals going through different
strings of coils that have their ends connected together. Why doesn't it all just short out and burn up?
That's where the rotations, and more importantly the phase angles, come in.
Check this out:
cos(0) = 1.0
cos(240) = -0.5
cos(120) = -0.5

1 + (-0.5) + (-0.5) = 0.
So what? Well, those are the cosine function values for the phase angles (shown in degrees) for our
previous three 3-phase equations. That they sum to zero is why our three-phase motor/generator doesn't
spark and smoke in normal operation. If you add the same fixed angle to the phase angles in our cosine
equations, the result is always the same. Try it:
cos(875634.43 + 0) = -0.4135812
cos(875634.43 + 240) = 0.9952783
cos(875634.43 + 120) = -0.5816971

-0.4135812 + 0.9952783 + (-0.5816971) = 0.


Try it again if you want to, but I guarantee the result will be the same. What this means is that no matter
what the rotational position of our phase sinusoids, if we look at them at the relative positions defined by
their phase angles, the electrical waveforms will sum to zero.
Now, by no coincidence, the connection locations for our motor/generator phase coil groups fall at the
relative phase angle positions. So, though the coil group ends are shorted together, each connection is
effectively at zero voltage, and nothing bad happens. You can see this relationship in, for example, figure
19. Consider the maximum up excursion of any signal in the figure to be +1, and the maximum down
excursion of any signal in the figure to be -1. Then, if you sum the observed values for the three sinusoids
at the same position along the horizontal axis, you will find the result is always zero. This relationship
will hold whether we are considering the rotation in degrees, frequency, or time. So long as we observe
the correct phase difference, the sum is always zero. And, of course, this extends to numbers of phases
other than three.
It might seem that connecting a motor/generator so that the voltages at its phase wires sum to zero won't
accomplish much. But, these are relative voltages. If you look across the phase coil groups instead of just
at each of the ends, you will see a voltage difference. This is why though the connection voltages sum to
zero, you should not go probing them with a wet finger! (Or a dry one, for that matter.)
The across phase-coil-group voltage difference relationship can also be seen in figure 19. Look at the
difference in voltage for each sinusoid over their specified 120 degree phase difference locations. That is,
from 0 to 120 degrees for phase A, from 120 to 240 degrees for phase B, and from 240 to 360 degrees for
phase C. In each case we see a difference of 1.5 (as a change from +1 to -0.5). The total excursion for
each sinusoid is 2.0 (as a change from +1 to -1), so we see across each phase 0.75 of the total excursion.
There are specific mathematical relationships for the continuous values of the phase voltage differences
over time. They are often expressed in terms of rotating vectors. We'll look at those as we investigate the
primary types of phase connection configurations.
You say delta, I say why?
Phasors don't have to stun:
Before we get on with looking at some of the possible motor/generator phase wiring configurations, first
lets take a look at a some notation to use for our rotating phases. For the most part, we aren't going to be
concerned with the absolute angles of rotation of our phases, but just with their relative phase angles, and
their magnitudes (which may be looked at in terms of either current or voltage). We know we are talking
about sinusoids all of the same frequency here, so, rather than using the cosine equations to formally state
that, we can just use a notation that gives us the magnitudes and phase angles we need. Here's one:
K

The parameter K is the magnitude of the sinusoid, and the parameter nnn gives the phase angle of the
sinusoid relative to the zero phase angle sinusoid. The phase angle parameter could be given in any units,
but, here, as indicated, we'll use degrees. So, you can read the notation as "K at an angle of nnn degrees."
This notation may look familiar. If you've done any AC circuit analysis it probably should. It is
traditional phasor notation as derived from the complex exponential representation of a sinusoid. We
aren't going to worry too much about its derivation here. Mainly we're just going to concern ourselves
with how to add phasors together to see if our motor/generator designs fit the criterion we established in
the last section, that the physical connection points for our phase coil groups are made where their phase
waveforms will sum to zero. For our purposes, often that will amount to just adding the phase angle
values to see if they sum to zero, where 0 in degrees is defined as some integer multiple of 360 degrees.

For example, ignoring the parameter K, (that is, call it equal to 1), we can sum the phasor equivalents of
our three-phase cosine equations as:
+

That is, our 3 three-phase phase-angles sum to zero. And, that summing to zero is the condition we want,
regardless of the number of phases we are working with.
Phase arrangements:
This section is brought to you by the letter Y. Why do people have to do all those the weird phonetic
things they do when they want to express the 25th letter of the English alphabet, Y? I'm dyslexic. I have a
hard enough time getting letters sorted out using a spell-checker, and all those things involving "w" and
"e" people do when they want to say "y" just make my head want to explode. So, here, when we talk
about "delta" systems, we'll call them delta systems, and when we talk about "Y" systems we'll call them
Y systems.
There are a lot of different ways to arrange the wiring of a polyphase motor/generator. But, they center
around the two basic configurations, the delta and the Y. In fact, most of the alternative configurations are
just combinations of delta and Y circuit arrangements.
Most of what we talked about so far has been while referring to delta wired motor/generators. But, the
discussion also applies to Y configured systems. The same phase coil group "head-to-tail" wiring method
described in the rote instruction sections holds for delta and Y arrangements, as does the notion of phase
voltages summing to zero at the phase group connection points.
What mainly differentiates delta and Y systems is the connection of their phase coil groups. We've seen
the delta connection before (figure 5), so, let's take a look at the Y connection. For our ongoing case of
eight magnets and 12 coils, we would connect our phase coil groups the same as we did for our delta
wired motor/generator (figure 4), but, rather than connect the phase coil groups in "head-to-tail" fashion,
we connect them "head-to-head." That is, if we consider the positive lead of a phase coil group to be the
"tail" of the first current arrow as drawn its coil connection diagram, and the negative pole to be the
"head" of the last current arrow in its coil connection diagram, then we connect the negative leads. For a
three-phase motor/generator, this gives us a three-armed or "Y" configured circuit, where the free ends
are our A,B,C phase input/output leads (figure 21).

Unless you look close at the wiring, the basic physical design we've seen for a delta wired
motor/generator (figure 6) looks the same for a Y configured motor/generator (figure 22).

So, what about that connection voltages summing to zero thing? Does that still hold for the Y
configuration? Yep. Take a look. Consider driving the Y as a motor. The free ends are provided with the
phase inputs, which we already know sum to zero due to their phase differences. Since the phase group
coils are identical, and we've connected them so that phase currents flow without interferring with
themselves, the voltages seen on the other ends of the phase coil groups will not change phase, and,
hence, still sum to zero at the Y connection point. Symmetry dictates the same holds true when using the
system as a generator. The same sinusoidal phase relationship diagrams we developed when discussing
the delta wired motor/generator also hold for the Y system.
Before we get into considerations of power and efficiency that might affect one's choice of a delta or a Y
configuration, there is one significant difference to note between the two connection styles. The Y
connection allows including a neutral line along with the phase lines if we choose to. By connecting a
lead to the phase coil group junction we get a three-phase 4-wire connection, while without the additional
lead the connection is known as a three-phase 3-wire connection (figure 23).

There are a number of reasons to choose to include or not include a neutral line in a polyphase AC
system. Mainly they relate to safety and power balancing considerations, and are not terribly relevant to
the discussion at hand. So, we won't cover the issue of neutral lines more in this section, or in any great
detail later.

One thing that is somewhat relevant to our current discussion is the matter of power and efficiency for the
delta versus a Y connected motor/generator. Way back near the beginning of all this I said something
about the delta being more efficient at lower rpms and that is why it was chosen as the configuration for
the initial polyphase rote design instructions. Some of you may disagree with that, if you know that for a
given rpm the Y produces a a higher voltage than the delta. Well, efficiency is in the eye of the beholder.
For a given rpm the delta produces more current than the Y, and, me, I'm looking for maximum current,
not maximum voltage, so the delta makes more sense for my purposes. But, lets take a closer look and see
about how to figure out if it makes sense for your application on not. Note that outside the standard
electrical current and voltage relationships which hold anywhere, the discussion that follows will refer to
three-phase systems only.
Whether we are looking at a delta or a Y system, there are two voltages to consider, coil voltage and line
voltage. Coil voltage is the voltage read across any coil, where "coil" means the entire phase coil group
when there is more than one coil per phase coil group. A motor/generator's A,B,C, phase leads are
commonly referred to as the "lines" and line voltage is the voltage read between any pair of lines (figure
24).

The following relationships hold for coil and line voltages and currents in a three-phase delta system:
Coil Voltage = Line Voltage
* Coil Current = Line Current
The following relationships hold for coil and line voltages and currents in a three phase Y system:
* Coil Voltage = Line Voltage
Coil Current = Line Current

Given those relationships, and the rule of thumb that the line resistance, (i.e., the resistance measured
between any two lines), in a Y system is typically about 3 times the line resistance in a delta system, we
can make a few calculations and look at differences in output from the two.
If we measure the voltage and current in one phase of a generator, (where here phase means one phase
coil group, or the "coil" as described in a previous paragraph), independent of the other phases, we can
use those readings to calculate what the line voltage and current will be for the generator connected in
either a Y or a delta configuration.
Say in one phase of a generator we see 28 volts and 10 amps. Then, using the above relationships, wired
as a delta the generator would have the same line voltage, 28 volts (V), and a line current of
* 10 =
17.3 amps (A). Rewired as a Y, the same generator would have a line voltage of
* 28 = 48.5 V, and
a line current of 10 A.
Now, the power in an electrical circuit in watts (W) is equal to the voltage in volts times the current in
amps, So from any two lines in our Y configured generator we get 48.5 V * 10 A = 485 W, and,
similarly, as a delta we see 28 V * 17.3 A = 484 W for any two lines.
Hmmmm, 485 W versus 484 W, that doesn't seem too significant a difference. And, it isn't. The power
out of a generator is related to the power you put into spinning it. Whether we wire it as a delta or a Y, if

we spin our generator so in either case we get the same phase measurements, then we are basically
putting the same amount of power into it. So, in that case, we really can't get any more power out of one
configuration over the other. OK, then why choose one over the other? It comes down what you are
trying to do, and how fast you can spin.
Say we made our above measurements at 400 rpm. For the delta that means 400 / 28 = 14.3 rpm per volt,
and for the Y we have 400 / 48.5 = 8.2 rpm per volt. If we reved up to 1000 rpm, from the delta we would
see around 1000 / 14.3 = 69.9 V, and from the Y we would see around 1000 / 8.2 = 122 V. Now, rather
than choose an arbitrarily coil current, lets use the convenient, but not unrealistic line resistances of 3
ohms for a Y system, and 1 ohm for a delta system and see what we get for power while we are charging
a battery.
Lets cheat a little bit and ignore the need for rectifying the output alternating current (AC) voltage to a
direct current (DC) voltage before applying it to the terminals of a battery for charging. As far as the two
types of voltages are concerned, rectification results in a known, fixed multiplication factor being applied
to the AC voltage level to obtain the DC voltage level, so the calculations that follow are valid, but just
missing a multiplier. We'll get to rectification in a later section.
The voltage (V) across a circuit is equal to the current (I) through the circuit times the resistance (R) of
the circuit, that is, V = I * R. We can manipulate that equation any way we like, and, hence, I = V / R,
which we use with our generator line voltage and resistance to calculate the line currents. Actually, to
calculate the generator line currents in the case of charging a battery, we subtract the battery voltage from
the previously calculated generator line voltages because the battery is compensating for part of the
measured line voltage. So, for a nominal 12.5 battery level charged by the Y configured generator
spinning at 1000 rpm we have 122 - 12.5 = 109.5 effective line voltage, which, with the line resistance of
3 ohms gives us 109.5 V / 3 ohm = 36.5 A line current. Similarly, for the delta we have (69.9 - 12.5) V / 1
ohm = 57.4 A line current.
A battery being charged by a generator is essentially a series circuit, and the current through a series
circuit doesn't change. (We'll talk more about series and parallel circuits later.) So, for comparison of the
Y and delta generator configurations, we can look at the power output relative to the nominal battery
voltage because we see the same current no matter where we look in the circuit, and it will give us an idea
of how much power we have available to charge the battery. That is for the Y we see 36.5 A * 12.5 V =
456 W, while for the delta we see 57.4 A * 12.5 V = 718 W, or 253 W more power available for charging
from the delta than the Y when both are spinning at 1000 rpm.
That more output power is available from the delta at a given rpm is an advantage so long as we have the
input power available to spin it fast enough to get the required output voltage. If we are constrained by
input power, and hence to lower rpm, then the Y configuration can be an advantage when, say, charging a
battery. To get the minimum required approximately 13.5 V for charging a standard lead-acid battery, our
example delta generator would need to spin 13.5 V * 14.3 rpm per volt = 193 rpm, while the Y would
require 13.5 V * 8.2 rpm per volt = 111 rpm. So, though we would not be producing as much power as
when spinning at 1000 rpm and hence would not be able to charge a battery as fast, we might still be able
to charge a battery with the Y configured generator when we could not with the delta generator due to
rpm constraints.
Some motor/generators take advantage of the voltage and current relationship difference between delta
and Y configurations and start in the Y configuration to take advantage of higher voltage and lower
current at low rpm, and then switch themselves to a delta configuration once the line current is sufficient
to energize a coil that creates a magnetic field strong enough to pull in a spring loaded switch that makes
the configuration change. (That's one way, anyway, there are other methods.) So long as the rpm stay
high enough to maintain the required voltage level, the motor/generator runs in the the delta configuration
and takes advantage of the higher available current. Otherwise the system drops back into the Y mode
and maintains the required voltage level at the lower rpm.
Rollin' Rollin' Rollin'
Well, rotatin' rotatin' rotatin' more precisely. Rotating magnetic fields, that is. Which are at the heart of all
we've been talking about. The notion of rotation we covered a lot already with our sinusoids and their
multiple forms of cosine equation descriptions. Plus, "moving" magnetic fields have been mentioned
before, too. It's pretty clear that the magnetic fields of magnets mounted on a spinning rotor must be
rotating, since their magnets are rotating. (Of course, there a few more details to fill in there.) And, what

about those magnetic fields created from our nonspinning armature coils when we apply phase input
power? Well, they are rotating, too!
In the last section we talked a lot in terms of generators. Here, we'll cover things for the most part in
terms of motors. Everything in this and the last section pretty much applies to motors or generators,
(remember symmetry), but, the phenomena are just easier to visualize as described. The key thing to note
is, like the term "phase," the term "rotation" can have somewhat different meanings depending on
context.
Moving? Rotating?
Moving a magnet obviously moves its magnetic field along with it. So, rotate a magnet, and its magnetic
field rotates, too. OK. That's a straight forward concept. But, let's look just a little bit closer.
First, we'll consider the magnetic field of a single magnet. Note that all magnets are dipoles, that is, they
have two components (north and south poles), that are inseparable. There are a number of formulae that
can come into play when looking at dipole magnetic fields. We don't need to worry about them too much
here, and just note that the strength of a magnetic field near a dipole falls off as 1/ , where r is the
distance from the center of the dipole. (This is different than the fall off from a monopole, such as a
positive or negative electric charge, where field strength falls off as 1/
.)
The 1/
relationship (figure 25) applies best on the molecular scale, but for a passable approximation it
will work to look at field strength from, for example, one of the neodymium discs on the model 3-phase
motor/generator rotor constructed for this web page.

If we ignore the mathematical artifact where the field blows up to infinity at a distance of r = 0, and kind
of fudge over the fact that our disc has a significant diameter relative to a molecule by smearing out the
field over the width of the disc and plotting the field change relative to the edge of the disc, we get
something that gives some idea of the field strength near the disc. The take home message here isn't the
exact form of the field, but that field strength falls off fairly rapidly away from the disc, and we are left
with essentially a magnetic field "bump" above our magnet. (figure 26).

Now, consider the magnets on the spinning rotor of our basic design motor/generator. What we have is a
series of intense magnetic bumps, with alternating poles, evenly spaced around the perimeter of the rotor.
Assuming the gap between the moving rotor magnets and the stationary armature coils is small enough
(we'll talk a bit about proper gap width later), then the "bumps" will move through the armature coils,
creating a current in the coils.
If we didn't have the magnetic bumps, but rather a uniform field around the rotor, then even though the
rotor was spinning, and the field effectively moving, the coils would not see much if any change in
magnetic field, and, hence, produce little to no current. It isn't just that a magnetic field is moving that
causes current flow in a wire, but that the field intensity is changing, and the moving magnetic bumps
give that necessary intensity change.
For our alternating pole rotor, we not only see relative level changes in field intensity as the rotor spins,
(i.e., a changing magnetic field), but we also see full alternating pole magnetic field changes. This
alternating of poles, as discussed before, causes reversal of (alternating) current in our armature coils.
Voila, AC generator!
In the case of using the motor/generator in motor mode, you can consider the magnetic "bumps" along the
edge of our permanent magnet rimmed rotor to be the equivalent of teeth on a gear. But, then, what is
pushing these "teeth" to make our motor rotor spin? The answer is the rotating magnetic field produced
by our armature coils when we apply power to its phase coils.
Inducing synchronization:
The brushless permanent magnet motor is what is referred to as a synchronous motor. That means its
rotor spins at synchronous speed, where synchronous speed is an rpm value that is a direct function of the
input phase line frequency and the number of armature poles in the motor. Under normal use conditions
there is no slip of the rotor in a synchronous motor relative to the rotating magnetic field created by the
phase line power input to its armature coils. This is as opposed to an induction motor, where there must
be some slip of the rotor (i.e., its speed less) relative to the rotating armature field so that currents are
induced in conductors embedded in the rotor which in turn create magnetic fields that act as the magnetic
fields produced by the permanent magnets on the rotor of our brushless permanent magnet motor. The
synchronous type motor makes a good motor/generator. The induction motor is not quite so well suited to
acting also as a generator.
Now, electrical current moving through a wire travels basically at the speed of light, which is about 1 foot
per nanosecond (0.000000001 second). And, since we are talking about input power frequencies
generally of less than 100 Hz, (giving a cycle period on the order of 0.01 second), with total wire lengths
on the order of tens to hundreds of feet, for all intents and purposes, every coil in a phase coil group sees
the same power level at the same time, tracking the input power level as it arrives at the phase line input.
So, if everything seems to happen at once, how do we produce anything like a rotating magnetic field
from our stationary armature coils? Well, there are a few aspects to that, in particular, the alternating
polarity wiring of the armature coils which produces opposite pole magnetic fields in adjacent coils as
current flows through them, and also, in the case of polyphase systems, the phase difference between the
phase line inputs which produces a lead or lag between magnetic fields produced in different phase coil
groups.

Probably the most commonly seen formula for synchronous speed is:
S = 120f/p
where: S = synchronous speed in rpm (revolutions per minute)
f = input phase line frequency in Hz (cycles per second)
p = number of armature poles (coils)

Let's stare at that just a tad and see if we can figure out what's going on. Since we are using frequency, f,
in Hz, or cycles per second, and S comes out in rpm (revolutions per minute) there must be a factor of 60
seconds per minute in there, so, we could rewrite the formula for S as:
S = (2)(60)f/p.

Now, if we wanted to convert from, S, speed in rpm to, say, N, speed in rps (revolutions per second), all
we have to do is toss out the factor of 60 seconds per minute, leaving us with:
N = 2f/p.
Apply a little algebra, and we get:
N = f/(p/2).

That is, synchronous speed in revolutions per second is equal to the input phase line frequency in Hz
divided by one-half the number of armature poles. That is to say, synchronous speed in revolutions per
second is equal to the input phase line frequency in Hz divided by the number of armature pole pairs.
OK, there's that "pole pair" thing again. So, where's it come from? Well, it comes from the same place it
did when we were looking at mechanical vs. electrical degrees a while back. There we had a sequence of
alternating pole permanent magnets. Here, due to the armature phase coil group wiring technique, we
have alternating pole electromagnets. In either case, to see the the same action at a pole we have to look
at every other pole. That is one half the poles give a different action, and one half the poles is also the
number of pole pairs. Hence, the number of pole pairs becomes an important factor in the calculations. In
fact, in a while we'll take a look at a diagram very similar to the one we looked at when discussing
electrical vs. mechanical degrees (figure 20) as we further define the concept of synchronous speed. But,
first, let's just take a closer look at the concept of an actual rotating magnetic field.
It is actually easier to conceptualize a rotating magnetic field when examining the properties of a
polyphase motor than it is when considering a single phase motor. In fact, a self-starting single phase
motor is, at least on startup, connected so that it actually runs as a two phase motor. (More on that to
come.)
In general, to establish a rotating magnetic field the number of armature poles (coils) must be equal to, or
a multiple of, the number of input power phases, with the poles being separated by the phase angle
between the input power phases.
As pointed out earlier, the number of armature poles and the number of rotor pole (magnet) pairs are
related. That is, for each rotor magnet pair must be a set of armature coils equal in number to the number
of input power phases. Note how this relationship, by default, meets the criterion given above for the
number of armature poles required to establish a rotating magnetic field.
So, we know how many armature coils we need to set up a rotating magnetic field, and that our properly
designed motor/generator will have that number of armature coils. Let's finally take a look at how all this
sinusoidal signal, spinning vector, rotating field stuff comes together. We can look a simple case, say a
two pole-pair, three phase, delta configured motor. Now, another diagram is in order. For that we can use
a new form of the armature coil connection diagram we've seen before, same as our original linear
diagrams, (e.g., figure 5), just wrapped in a circle to represent a more realistic motor/generator armature.
For our two pole-pair, three phase motor/generator we will end up with a circular figure having six
armature coils representing two three-coil phase groups (figure 27).

Using figure 19 as a reference for the phase relationships of three-phase sinusoidal currents, let's examine
the magnetic fields that would be generated in the coils of figure 27 by application of three-phase power.
Consider that a positive current through a coil wired as is coil A1 relative to its input phase power will
produce a north magnetic pole pointing inwards towards the center of the ring of armature coils.
We've seen that we can look at our sinusoidal signals in terms of degrees or time and get the same results.
So, take time zero, to be the zero degree position for phase A, take time two, to be the zero for
phase B, (120 degrees for phase A), and take time four,
to be the zero degree position for phase B,
(240 degrees for phase A). Note, relative to the waveforms in figure 19, this means at time phase A is
at a positive maximum, at time phase B is at a positive maximum, and at time
phase C is at a
positive maximum. But, let's concentrate just on phase A for a moment.
Note while coil A1 produces a maximum strength north pole oriented towards the center of the armature
ring at coil A2, being wired for the opposite polarity, produces a maximum strength south pole
oriented towards the center of the armature ring. The physical arrangement of coils in figure 27 puts A1
and A2 opposite each other on the armature ring, which means at time there is the maximum possible
magnetic field strength between coils A1 and A2. Also note that at the 90 degree position for phase A, or
three quarters of the time between and the A phase value drops to zero, and at that time there is no
phase A current generated magnetic field between coils A1 and A2.
Of course, there are two other input phases, and, as we already know, none of the three are always at their
maximum value. Again referring to figure 19, we can see that at time when phase A is at its positive
maximum value, both input phases B and C are are at one half their maximum negative value. Similarly,
at time phase B is at its positive maximum while phases A and C are at one half their maximum
negative value. And, at time phase C is at its positive maximum with phases A and B at one half their
negative maximum. For now let's just concentrate on time
We have already established that at time the magnetic field between coils A1 and A2 is at its
maximum, and oriented north to south from A1 to A2. For phase B at recall the rote coil connection
scheme dictates that coil B2 is the phase input coil and thus wired so that a positive current generates a
magnetic field with its north poll oriented towards the center of the armature ring while coil B1 is wired
so a positive current generates a magnetic field with its south pole oriented towards the center of the

armature ring. So, with the B phase being negative at coil B2 generates a magnetic field with its south
pole oriented towards the center of the armature ring and coil B1 generates a magnetic field with its north
pole oriented towards the center of the armature ring, and develops a total field strength around one half
that of the field between coils A1 and A2. Similarly, the total field between coils C1 and C2 is
approximately one half the field between coils A1 and A2, oriented north pole to south pole from coil C2
to coil C1. Well, at least that is what would happen if the different phase fields didn't interact. In reality,
the full strength field extends between coils A1 and A2 as described, but, the weaker north pole field
from C2 is deflected by the stronger north pole field from coil A1 and connects to the weaker south pole
field from coil B2, while the weaker north pole field from coil B1 is deflected by the field from coil A1
and connects to the weaker south pole field from coil C1. In a simplified form, the arrangement of flux
lines in the armature ring at time looks like the depiction in figure 28.

We could go on producing diagrams similar to figure 28 to show the orientation of magnetic flux lines in
the armature ring at times other than But, that would get very tedious very fast. Instead, we'll switch to
phasors and rotating vectors to simplify things.
Getting into the spin:
A few sections back we introduced basic phasor notation. There we saw how it describes a vector by
giving the vector's magnitude and direction. Lets dig into it a bit deeper and look at how it can represent
spinning vectors.
Phasor notation gives the direction for a vector as an angle relative to a fixed starting point, the tail end of
the vector. Since one end of the vector is fixed in place, if we change the direction angle, the vector will
pivot around the fixed end point until it aligns with the new direction angle. And, if we change the angle
value continuously in one direction the vector will spin around its fixed end point (figure 29).

OK, it's clear that phasor notation can represent a rotation. But, how to make that rotation represent our
motor/generator phases? Well, all we need do is force the angle argument to change in time with the same

cycle rate as the motor/generator input phases. And, it's radians to the rescue! (Again.) Just using the
same angle arguments we previously developed for our cosine equations as the angle argument in the
phasor notation vector representation gives us vectors spinning with the same cycle rate (frequency) and
phase differences as our motor/generator input phases.
If we view the rotating vectors (phasors) as something akin to the simple machine presented in figures 15
and 16 for converting circular to linear motion, then the relationships for the phasor magnitudes can be
deduced. Since our input phases are identical except for their phase differences we need look at just one
of them to determine the magnitude relationships for all. (That relationship is, of course, to scale the
maximum magnitude by the cosine function. But, for the sake of drawing some more pretty pictures, let's
soldier on a while longer.)
Looking carefully at how the our conversion machine works, in figure 15 we can see that the position of
the peg on the disc determines the amplitude of the waveform the machine produces. At any time in a
revolution of the disc, regardless of the angle of the peg relative to the machine's axle shaft, the
perpendicular distance of the peg from a horizontal line drawn through the center of the axle is equal to
the amplitude of waveform traced by the pencil attached to the end of the shaft moved by the peg.
If we think of the peg in our machine's disc as marking the head of a rotating vector (phasor) and the axle
shaft as marking the fixed tail point of that vector, then we can also think of the perpendicular height of
the tip of our rotating vector relative to a horizontal line drawn through its fixed tail end point as defining
the amplitude of the waveform drawn by our machine (figure 30).

Clicking on the thumbnail below will launch an animation demonstrating the change in amplitude with
phasor rotation. The animation cycles 25 times. If you haven't gotten bored with it before then and
already hit the back button in your browser to return to this text, then you can hit your browser's reload
button to run the animation again, until you do get bored.

We have established that our machine drawn waveform describes a sinusoid, so, our rotating vector also
describes the same sinusoid. This means our rotating vector (phasor) sinusoid can be used to define our
motor/generator phase input sinusoids, just as we have already done using the cosine functions that
describe the waveforms produced by our circular-to-linear motion conversion machine. We have also
shown how to graphically produce the proper magnitude for our phasors. But, we aren't quite done yet.
We need to apply both the direction and magnitude simultaneously to properly illustrate the phasor in
action.
By definition, a vector, rotating or not, describes a length (magnitude) and the direction to point that
length towards. To complete our picture, we need to not just project the magnitude from tip of the vector
onto a line as done in the previous figures, but, take the projected magnitude values and align them in the
direction of the vector at the time they were generated.
Considering again just phase-A, with no phase angle, and clockwise rotation. The phasor starts out with
its positive maximum magnitude directed vertically up at an angle of 0 degrees. The only other maximum
is negative occuring when the vector is again vertical, only directed down at an angle of 180 degrees. At

90 degrees and 270 degrees the magnitude is 0. The effect is in one cycle for the phasor to scribe a figureeight shaped path around its fixed tail point (figure 31).

Clicking the thumbnail below brings up an animation that illustrates the change in magnitude with
rotation describing a figure-eight shaped path.

Now, don't be fooled. The phasor described above isn't really our long sought after rotating magnetic
field. The phasor, in this case, simply describes the change over time in the magnetic field relative to the
phase A coils, A1 and A2. Between coils A1 and A2 the magnitude of the magnetic field generated due to
the phase A input increases and decreases, as well as changes its polarity when the input current changes
polarity, but, in and of itself, it doesn't rotate. The rotation indicated in the figures and animations above
is just really a way of illustrating the change over time. None the less, the "figure-eight" magnitude
change is the key to the rotating magnetic field.
We noted in describing the magnetic flux diagram of figure 28 that, because of the way the phase coil
groups are wired, at time the fields from each group are directed towards the center of the armature
coil with the same magnetic polarity, north to south, but have different magnitudes, with the strongest
flux associated with the phase A coil group. No matter at what time in an input phase cycle we look, the
flux across each coil group will always have the same polarity, though all polarities may be reversed and
which phase coil group is exhibiting the maximum flux will change. It is this shift in flux magnitude
between the phase coil groups which leads to a rotating magnetic field due to the polyphase input signals.
For our three-phase system we have three rotating phasors, each leading the next by a phase angle of 120
degrees. If we plot the figure-eight magnitude changes for each phasor on the same diagram, the result
provides more insight into the nature of the rotating magnetic field. From such a diagram (figure 32) we
can see loops in the magnitude change plots for the individual phasors intersect at the center of the figure,
but each fills the gap between its adjacent phase loops, with some overlap. As explained in the text of
figure 32 this figure can be used to determine the magnitude and direction of the phase coil magnetic
fields in our 3-phase motor/generator.

Still, we haven't quite got to the real deal rotating magnetic field. The animation which can be viewed by
clicking the thumbnail below may help shed some light the matter. It shows the figure-eight sweeping
magnitude for all three phases, in their proper orientations, at the same time.

Again, recognize that in the above figures and animations we are not looking at rotations in the sense of a
magnetic field moving around the armature ring. What we are looking at is the change over time of the
flux associated with each phase coil group. The apparent rotation in the figures is just a way of
representing the repeating cycle of the changes. It is the combination of the changes in all phases at once
that produces the rotating magnetic field.

It all adds up:


OK. We have our rotating vectors, also called phasors, spinning in sync with our phase inputs, and having
magnitudes representing the current flows through our phase coil groups which are proportional to the
magnetic field fluxes from the armature phase coils. Also, a while back, we noted at time (figure 28)
that because of how the phase coil groups are wired the polarity of the magnetic fields from each group is
oriented in the same direction towards the center of the armature ring. Again referring to figure 19, let's
call the negative maximum for phase C time t1, the negative maximum for phase A time t3, and the
negative maximum for phase B time t5. Similarly to how we did for time to produce figure 28, we can
take a look at the fluxes for the phase coils at time . Except this time instead of producing another
picture like figure 28 we'll use vectors to make our graphic representation.
At time , the picture is simply the reverse of the picture at time . The magnitudes of phases B and C
are one half the magnitude of phase A, with phase A current flow negative and phases B and C current
flow positive. As we saw when discussing figure 28, due to the wiring of the phase coil groups polarities
of the magnetic fields for all groups are oriented in the same direction towards the center of the armature
ring, but this time instead of north to south from coil A1 to A2, it is south to north from coil A1 to A2,
with weaker south to north fields from coil B1 to C1 and from coil C2 to B2. To be consistent, from here

on we will always talk of our magnetic fields in terms of flux from north to south. So, at time t3 we will
say our fields are north to south from coils A2 to A1, C1 to B1, and B2 to C2.
Now, how do we represent the situation at time or with vectors? Pretty straight forward. We know
the lengths of our vectors represent magnitude, so, at time or the phase A vector will be twice the
length of the phase B and C vectors. Further, we know the orientation and wiring of our phase coils, so
we can determine the polarity of our phase coil magnetic fields and hence the direction to draw our vector
arrow heads. This means at time , from the position of coil A1 on our armature ring we draw a full
length vector representing the phase A magnetic flux to coil A2, and, from the same starting point two
half length vectors, one at 60 degrees from the right of the full length vector and the other at 60 degrees
from the left of the full length vector. Examining figure 28, we can see that the half length vector pointing
60 degrees to the right represents the B phase north to south contribution to the picture (coils B1 to C1)
and the half length vector pointing 60 degrees to the left represents the C phase north to south
contribution (figure 33A). At time the full length vector (phase A contribution) points, north to south,
from the coil A2 position with the phase B contribution (coil B2 to C2) at 60 degrees to the left, and the
phase C contribution (coil C1 to B1) at 60 degrees to the right (figure 33B).

But, "wait!," you say. Aren't the phase separations 120 degrees? Where did those angles of 60 degrees
come from? Well, that's really just a bit of geometric handwaving coupled with the knowledge of how our
armature phase coil groups are wired. In discussing figure 28, we noted how the B and C phase field
contributions are polarized due to their phase group coil wiring. A little squinting at figure 28 will reveal
that relative to the center of the armature ring, directly at the edge of the armature ring, the contributions
of the B and C phases point at 60 degrees relative to the phase A component. The fact that the sum of the
angles of the B and C field contributions sum to 120 degrees is really just a coincidence, and is related to
the number of coils on the armature ring. A different number of coils in the phase groups would give a
similar picture, but the angles would be different. The important thing to remember here is the phase
input vectors and the magnetic field vectors are, of course, strongly related, but not the same. The phase
input provides current flow, and current flow through the wire creates the field.
Note that the picture at time is simply a version of the picture at time rotated by 180 degrees about
the center of the armature ring. Not coincidentally, 180 degrees is the amount of phase rotation we see
from time to time . That is, when our input phases have rotated 180 degrees, it appears that so do
our magnetic field flux vectors. We can produce the same kind of figures for times and
(phase C
maximums), and times and
(phase B maximums), where the diagrams for times and
will be

versions of the and


will be versions of the

diagrams rotated 120 degrees clockwise, and the diagrams for times
and diagrams rotated clockwise 240 degrees (figure 34).

and

One more thing to note is we are referencing our rotation to the center of the armature ring, and what we
really have is a south magnetic field on one side of the ring, and a north magnetic field on the opposite
side of the ring, The magnetic field rotation is of both the poles. We have been drawing our diagrams to
reference the south magnetic part of the field, but the north magnetic field part rotates along with it, in
sync, around the opposite side of the armature ring.
At this point we basically have our rotating magnetic field. At least we can see at times , , , ,
, and we have the same magnetic field flux pattern rotated at 60 degree increments around the
armature ring, where 60 degrees is the amount of phase rotation seen between each time step.
Another important thing to note in figure 34 is that it is not the flux vector for phase A that is rotating
around the armature ring. You can see in the figures for each time step that the main flux comes from
which ever phase is at its positive or negative maximum value. If you compare the phase group coil
layout around the armature ring of figure 28 with the time step diagrams in figure 34, you can see that the
spacing of the coils results in the equal spacing of the direction of the flux vectors around the armature
ring. Since we can't really tell one flux vector from another, what we see is an apparent rotation of a
single flux vector with time.
For the complete picture we would need to look at what happens between the points of phase magnitude
maximums. But, for this discussion, that is really more than we need to do here. The bottom line is as the
phase inputs vary for times away from the phase input maximums, the fluxes from the phase group coils
will sum up as vectors to create a total flux picture which looks basically like what we see for the phase
magnitude maximums, but with the head and tails for the main sum vector directed somewhere between
the phase group coils, rather than directly through an opposing set of coils. (Remember, the diagrams we
have seen are simple representations of a more complex picture.) In the end we have a magnetic field flux
vectors generated by the phase coil groups which vary so we end up with a set of north and south pole
fields moving around the armature ring such that they appear to smoothly rotate.
Whew! Finally! A rotating magnetic field. Wow, that was kind of like pulling teeth, wasn't it? Well, as
Blaise Pascal once said, "I made this letter long because I didn't have time to make it shorter." I may
clean this up a bit someday. But, don't hold your breath. For now, lets continue on the trail of that oft
promised, but yet to be seen self-starting single-phase motor.
Finding Another Way:

The final key to understanding the self-starting single-phase motor is seeing why a polyphase motor spins
in a given direction, and how it can be made to spin in the opposite direction. The direction of rotation is
determined by the sequence order of the input phases. It will also be helpful to understand a bit more of
the possible wiring configurations for a motor/generator. So, first will take a look at series and parallel
wiring arrangements, then we'll get on with the matter of direction of rotation.
Series and parallel and such:
We've hinted at other possibilities for armature coil phase group wiring configurations before, noting that
wiring coils end-to-end, with proper consideration of phase angles, allows addition of the voltages
generated in each coil. This means that besides putting more turns in our coils to generate higher voltates,
which can make motor/generator construction problematic due to large coil sizes, we can also use smaller
coils wired end-to-end (formally called a series connection) to achieve higher voltage outputs. For a
single-phase motor/generator such as the one illustrated in figure 3 this is simply a matter of adding more
coils to the armature ring, and an equal number of magnets added to the rotor disc. In the case of
polyphase motor/generators like the ones depicted in figures 6 and 22, to increase output voltage we
would add equal numbers of coils to each phase coil group, and add magnets on the rotor disc following
the constraints given by the design equations provided in the rote formula discussion.

LEAD ACID BATTERIES


How is it made and how does it work ?
The lead acid battery is made up of a series of identical cells each containing sets of positive and negative
plates.
In semi-traction cells, flat plate construction is used. Each positive plate is a cast metallic frame which
contains the lead dioxide active material. The negative plates contain spongy lead active material. also on
a similar frame. Both plates usually have the same surface areas.
In practice a typical cell is constructed with many more plates than just two in order to get the required
current output. All positive plates are connected together as are all the negatives. Because each positive

plate is always positioned between two negative plates, there is always one or more negative plate than
positives.
The resultant voltage of lead acid cell is normally 2 volts In order to achieve the voltage required for the
application each cell is then connected in series by substantial metal straps to form a battery. In a typical
monoblock battery, such as that used in a car for starting, the voltage required is 12 volts, achieved by
connecting six cells together in series and enclosing them in one plastic box.
Leisure batteries where a sustained current requirement is needed and a deep cycle, the ability to be
discharged to 90%, have a different make-up to that of a traction battery that is used in a car.
The cell containing the plates is filled with an electrolyte made up of sulphuric acid and distilled water
with a specific gravity of 1.270 at 60deg F (15.6deg C). Sulphuric acid is a very active compound of
hydrogen and sulphur and oxygen atoms. Sulphuric acid is a very reactive substance and because of its
instability it is able to distribute itself very evenly throughout the electrolyte in the battery.
Over time, this action ensures that an even reaction can occur between all the plates. producing voltage
and current. The chemical reaction between constituent parts of the electrolyte and the 1. spongy lead of
the negative plates and 2. the lead dioxide at the positive plates turns the surface of both plates into lead
sulphate. As this process occurs the hydrogen within the acid reacts with the oxygen within the lead
dioxide to form water.
The net result of all this reaction is that the positive plate gives up electrons and the negative plate gains
them in equal numbers, thereby creating a potential difference between the two plates. The duration of the
reactions producing the cell voltage is limited if there is no connection between the two plates and the
voltage will remain constant.
If a connection (a load) is placed between the positive and negative plates the chemical reaction is able to
continue with electrons flowing through the circuit from the negative plate to the positive. The flow of
electrons is in fact the current produced by the cell. Only when the supply of electrons becomes depleted
i.e. when the active material on the negative plate has been used up, and the within the electrolyte has
mostly been turned into water will the battery fail to produce any current. During the chemical process
different levels of heating can occur and the faster a battery is exhausted the greater will be the heating
and thus the efficiency of the system will be reduced.

Care and Maintenance of Flooded Lead Acid Batteries


The most important aspects of care for these and all other types of batteries concern both charge and
discharge as well as the mechanical treatment of the batteries i.e. keeping them topped up with water etc.
It is important to consider that lead acid (pb) batteries are, as we have seen above, quite delicate chemical
factories.

Discharge
Lead acid batteries should never be run flat. The maximum recommended discharge is 75% of the total.
This means that the battery should have a minimum of 25% of charge remaining when it is put on charge.
Lead acid batteries once filled with electrolyte, should always be regularly charged even if they are not in
use. When left idle a filled battery will self discharge because of its own internal resistance. left long
enough a battery can go completely flat without ever having been put into service.
Storage also affects the rate of discharge. A battery should never be stored directly on the ground and
especially not on concrete. The best storage method is wooden pallets which do not conduct or allow
damp paths and do allow good air circulation. During storage, most manufacturers recommend a
freshening charge once every two months or so.

How to find the state of charge


To easily see the state of charge accurately you should obtain and fit a battery condition instrument. The
most accurate way of checking a batterys condition is with a drop tester but this is not a quick and easy
method.*
(* These are available from various suppliers at reasonable prices in the UK. - Ed)
The cost of a battery condition meter or E meter may initially appear prohibitive, but you should consider
how much its cost will save by preventing the premature failure of expensive batteries.
It is extremely difficult to accurately measure the state of charge of a lead acid battery and to predict the
remaining capacity.

Battery capacity is not comparable to a tank full of petrol. A filled petrol tank contains a finite amount of
energy which can be used either slowly or fast according to the energy required. Battery capacity is not so
simple.
In a battery, the rate of which energy is drawn affects the overall amount of energy available from the
battery. For example , a 100 Ah battery rated at 20 hour rate means that over 20 hours there are 100 Ah
available, or to put it another way you can expect to draw up to 5 Amps per hour for up to 20 hours,
20x5=100Ah.
If you try to draw the battery down more quickly you cannot expect to get the same amount of Ah from it.
For example if you draw it down over just one hour the approximate capacity will become 100 x 0.59
which =59Ah. Putting this another way means that you can connect a load on the battery which will draw
59 Amps for just one hour. If you discharge over just half an hour you can only expect to get around half
(47%) the capacity from your batteries.
With most uses the rate at which a battery is discharged varies enormously , you can see that any battery
condition indicator has to be quite a clever piece of equipment if it is ever going to get close to giving an
accurate reading.
The E meter samples the rate of discharge every 4 minutes or so (this can be varied) and recalculates the
amount of time remaining before the battery will be fully discharged and also up-dates and displays a
fuel gauge bar graph. The combination of the fuel gauge and the time remaining displays along with
possible displays of volts, amps.
Ah and/or kWh means that the instrument is around 99% accurate and can be relied upon. There is one
proviso though, the E-meter must be correctly set up in the first place for the capacity and the type of
batteries you are using. It does take a bit of fine tuning to get it set up correctly, but once it is, there is no
better instrument that I have come across.
If an E-meter is not available a voltmeter giving the open circuit voltage of the battery can give an
approximate indication of battery discharge. This method cannot be relied upon as this voltage will rise if
the battery is allowed to rest and then the voltmeter will in effect give a false reading. As soon as the load
is reapplied to the battery the voltage will drop and the available charge will then become apparent.

Charging Lead Acid Batteries


Battery vent caps should always be kept in place and tight during both charging and discharging.
For best battery life, i.e. greatest number of charge/discharge cycles and years service most battery
manufacturers recommend that you should aim to recharge the batteries when they have reached around
50% discharge. This level of discharge, of course, must be measured according to the rate at which the
batter is discharged, which as w have already seen, varies the available total capacity of the battery.
In addition, some battery manufacturers specify that best life will be achieved only if the batteries are
discharged sufficiently for a 4 hour bulk charge to take place before the batteries are fully recharged.
A reasonable rule of thumb is that you should aim to charge the batteries only when they are between
70% and 40% discharged. If you charge them then they are only lightly discharged i.e. less than 40% you
will end up boiling them unnecessarily which wastes energy in the form of heat and gassed off hydrogen
and in turn shortens the life of the batteries.
In effect the batteries are being overcharged which can cause degradation and buckling of the plates. In
the process some active material is forced off the plates and drops down to the bottom of the battery. If
this occurs frequently the eventual result is a build up of a bridge between the plates which in turn can
cause a possible short across the plates. This situation leads to the destruction of a cell which then reduces
the capacity of the battery.
To confuse matters further, a battery will operate at its most efficient the deeper it is discharged, up to
around 75%. The bulk phase of the charge cycle is the most efficient and is proportionately longer the
deeper the discharge.
There are many battery chargers available. Today, manufacturers supply automatic chargers which are
supposed to ensure optimum battery life. there are numerous charge profiles available and now, with the
advent of the electronic switch mode chargers, it is possible to have a fully programmable charger on
board your vehicle capable of charging almost any type of battery.
Lead acid batteries must be charged carefully. If the charge is too violent and uncontrolled the batteries
can overheat and cause thermal run-away which can result in a possible explosion.

Too gently charging will take too long to get the batteries fully charged with the result that the batteries
will end up being used in an under charged state eventually leading to premature failure due to
sulphation.
The latest electronic chargers mostly make us the IUI charge profile for standard flooded lead acid
batteries. This means that the current drawn by the batteries is allowed to flow at a constant (I) rate while
the voltage is allowed to rise of its own accord, which it will do as the battery starts to be charged up.
This first part of the charging cycle is known as the bulk charge phase.
When a preset voltage has been reached, normally the voltage at which the batteries just start to gas, the
charger will switch into the constant voltage (U) phase and the current drawn by the battery will gradually
drop until it hits another preset level. This second part of the cycle is really the finish charge where the
battery just topped up to the brim very carefully at a much gently diminishing rate. Finally the charger
will switch again into the constant current mode (I) and the voltage is allowed to rise again, up to a new
higher preset limit, in order to achieve a successful equalization charge.

The Bulk Charge


In this first part of the charge the battery is allowed to have a large draw on the available current. Usually
the limit to this current level is determined by the availability of a suitably sized mains outlet, especially
on large batteries. It is however, worth noting that that the life of a battery will be greatest if even this
first bulk phase of charging is started off gently and the maximum current is limited. If the current is too
high the result will be excess heating within the battery which is wasteful and could lead to buckling of
the plates and destruction of the battery. Sizing of the charger to suit the batteries is important.

Finish charge
Once the bulk phase has been completed, the finish phase commences and the battery charge is topped
off. This phase is very important. If the battery is not topped up gently it will overflow in the form of
waste heat and violent gassing of the plates which again can lead to the plates buckling and the battery
being destroyed. If the battery is not topped up fully, it will become sulphated after only a few charges
and the result will be premature failure.

Equalization
In any cyclic application, a series of batteries will always need to be equalized from time to time in order
to ensure that the battery cells remain at the same voltage throughout the pack.
No two battery cells or batteries are created equal. During both charge and discharge each and every
cell/battery will react in a minutely different way to its neighbour. This could mean that each battery may
be holding a different quantity of charge. In order to get the most out of the total battery pack it is
necessary to make sure, as far as possible, that each and every battery is holding a similar amount of
charge.
During the charge cycle the voltages of the different batteries will very. In order to bring them all to the
same level it is necessary to give some a slight overcharge in order to bring the other up to full charge.
Equalization is done by allowing the voltage to rise while allowing a small constant current to the
batteries. The voltage is allowed to rise above the normal finish voltage in order to allow the weaker
batteries/cells to draw more current. The stronger batteries will not be adversely affected providing the
current is gently and the period and frequency of overcharging are not too high and great respectively.
The stronger batteries will absorb the overcharge by giving off heat by gently boiling and gassing more
heavily. Once the weaker batteries have absorbed the required current, the equalization charge can be
halted. The equalization time should be long enough to bring all the batteries up to a full state of charge.
As the time factor will very the most reliable way to check the charge states is by a voltmeter on each cell
or individual battery.
Really sophisticated battery charging and monitoring systems do not require the use of an equalization
charge and are able to charge all the batteries fully including the weaker ones without overcharging the
strong ones.
In these systems, each battery is fitted with an electronic clamp, which gradually reduces the amount of
charge going into the fully charged batteries as the finish charge progresses. This means that the weaker
batteries receive more current to bring them up to a full state of charge and the strong batteries are
prevented from being overcharged unnecessarily. The drawback with these sophisticated systems is their
cost. The price of each battery clamp can be in the order of 1/5 the cost of each battery.

Watering
Traction and semi traction batteries are generally supplied with removable vent caps so that they can be
kept topped up with water. The action of charging the batteries causes gassing when a certain voltage is
reached, usually somewhere around 2.35 to 2.4 volts per cell. The result is that water is depleted of its
constituent parts by liberation of the Hydrogen gas plates.
Hydrogen is of course much less dense that air and the electrolyte and consequently floats out of the
batteries at the earliest opportunity. This water must therefore be replenished from time to time. If you do
not gas a battery the chances are that it will become sulphated due to fact that not all the sulphate will be
fully removed during the non gassing phase of charging.
It is very important to note that only very pure water is suitable for topping up. It must contain no mineral
traces and especially no metallic solids, especially iron.
The most suitable form of water is distilled water or water that has been chemically treated and
demineralised. Only these types of water should ever be used to top up the batteries.
Tap water will quickly corrode the battery and should never be used.
The frequency of topping up required for the batteries depends on how they are used. If the are frequently
and heavily discharged they will need to be topped up regularly. Perhaps every two weeks to a month.
If the batteries are not charged and discharged frequently then it is likely that they will not require
topping up for longer periods of time, say once every two or three months or so.
If the batteries are regularly charged after only short discharges they will use much more water than
normally. This type of treatment of the batteries should be avoided. It is most important to make sure that
the tops of the plates in a battery never become exposed. They should always be covered by electrolyte or
they will quickly sulphate and the battery will fail.
It is possible to fit either automatic watering systems or catalytic caps. With the former, the battery vent
caps are replaced with special caps, which incorporate float valves.
A series of hoses connect the float caps to a single filling point. The filling point can be arranged so that
distilled water is poured into it from time to time or a reservoir can be fitted to constantly top up the
batteries. The catalyst caps prevent the battery charging gases from venting to atmosphere and by
chemical action combine the gases back into water. Thus there is little or no water loss.
Both the above systems work but there are drawbacks. Lots of hoses interconnecting battery cells and
batteries can pose a serious safety threat if one of the cells goes into thermal run-away and the vent caps
block and the tubes have no water in them. The result is that ignition from one cell will almost
instantaneously result in ignition in all the rest of the cells as the tubes will fill very quickly with
hydrogen gas which is extremely combustible and explosive, i.e. like a bomb!
A study by Lucas back in the 1970s showed what can happen. In a demonstration, two batteries were
connected via their vent float caps by a 3-mm bore tube of a length of over 60-ft. Combustion was
artificially initiated in one battery and within an almost immeasurably short time combustion occurred in
the second battery.
The moral of this tale is that only vent float caps in first class condition should be used and at around 500 to 8-00 each they are not cheap! In addition, should the tubes ever become blocked or the floats fail
to operate correctly, it is highly likely that the batteries will not receive the correct supply of water and
the result will be failure of at least one cell.
The catalyst caps are possibly a better bet. They are however, even more expensive and do not last for
ever. Most sealed batteries make use of a built in catalyst system.
The cheapest method of watering is by hand. It is also the most reliable provided it is done regularly.

Temperature
Temperature affects charging as well as discharge. As a rough rule of thumb, the cooler the temperature
the more charge a battery will absorb and the warmer the temperature the more it will discharge.
There are of course limits. However it is senseless to charge a battery that has just been discharged fast
and has been left out in the sun. In these circumstances it must be allowed to cool before charging is
initiated. The ideal charging temperature varies but a reasonable guide is between 15 and 25 degrees
Centigrade. Ideal discharge temperatures are between 20 and 30 degrees centigrade.
Unfortunately discharging a battery at a higher temperature, while desirably affecting its performance,
will adversely affect its life. In fact, so far as I can determine almost every battery manufacturer can work

out the number of Ah which can be drawn from any battery over its life under given constant discharge
conditions.
There does appear to be a finite number of Ah that can be drawn from a battery for a given discharge
profile. Predicting this discharge profile is the difficult bit!

Battery Care Check List


1. Keep battery clean and dry - dampness lets electric current leak away.
2. Keep vent plugs in place to stop dirt falling into cells.
3. A thin coating of petroleum jelly helps prevent corrosion of terminals and connections.
4. For topping up the cells, use either distilled water or clean rainwater preferably collected in glass or
plastic. Never top up the battery with anything other than distilled water or rainwater. Do not top up
battery with acid, unless on the advice of a battery technician.
5. Make sure that the positive and negative plates inside the battery are covered with electrolyte at all
times. Do not overfill.
6. Avoid adding water to a battery just prior to taking a SG reading, as the reading will be misleading. If
water has to be added, the battery should be charged for a while to mix it with the electrolyte thoroughly
before the reading is taken.
Maintenance Schedule:
(i) Check SG of electrolyte:- 1 month
ii) Check level of electrolyte top up if necessary:- 1 month
(iii) After boost charge, check cell voltages.
These should correspond to each other to within 0.05 volts:- 1 - 6 months
(iv) Check tightness of terminals and remove corrosion if necessary:- 6 months
DO NOT: top up battery cell with water when the battery is in a state of discharge. If the electrolyte level
is very low, top up only to make sure the plates are covered and no more. The fluid level rises with the
charge level, so if water is added when the batty is discharged, it may overflow on charging and lose
electrolyte.
DO NOT: "tap" into part of your battery bank to obtain lower voltages for running lower voltage
appliances. You will damage the battery bank by discharging some cells in relation to the rest of the
battery bank.
DO NOT: lift batteries by the lugs or terminals. Batteries need to be adequately supported from
underneath.
DO NOT: go near the batteries with an open flame or cigarette. You may cause the batteries to explode.
DO NOT: overcharge your battery bank to point of heating the cells up. This will cause inter-cell
damage. It is acceptable to charge, to the point of electrolyte bubbling. You may need, to add water if the
electrolyte level goes down.
DO NOT: install batteries in parallel if it can be avoided. To increase battery capacity you should
endeavor to get a single bank of the required amp- hour capacity rather than smaller batteries hooked up
in parallel. For example, six 2 volt cells connector series to provide 500 amp-hours of capacity is
preferable to two 12 volt, 250 amp hour batteries connected in parallel.
Batteries in parallel should either be protected or electrically isolated through the use of diodes or fuses.
This will ensure that if one battery fails due to a shorted cell, the current rushing from the good battery to
the defective battery does not overload the conductor risking a fire. It will also save the charge and
perhaps the life of the sound battery.
DO NOT: use alligator clips or other sprung jaw methods if possible, as sparking often occurs when they
are removed or attached. Hydrogen gas is generated by batteries under charge which is very explosive in
the presence of air. Sparking can ignite it. The resulting explosion will not only destroy the battery but
also injure the person holding the alligator clips with flying debris and battery acid.

Battery Installation
Heed the following points when installing a battery bank:

1. Lead Acid Batteries should be installed in a cool well ventilated area, well away from any source of
heat and from windows admitting direct sunlight.
2. Open stands should allow access from both sides for maintenance and cleaning.
3. Always keep cells upright to avoid damage or displacement of plate assemblies.
4. Never lift cells by the terminal lugs; large cells may be lifted by their handles (if fitted) or by means of
a sling made of plastic sheeting.
5. Cells must be placed on a flat surface for even weight distribution, and should never be rested on the
edges of packing cases etc.
6. Levers of any kind must not be used to position cells, instead a cell must be lifted bodily and lowered
gently. into position.
7. Never slide a battery across a floor; this particularly applies to those with acrylic cases.
8. When batteries are installed in cabinets, adequate ventilation must be provided to avoid a dangerous
concentration of hydrogen. Cabinet doors should be open during gas charging.
9. Stands should provide support for at least 50% of the base area. It is recommended that timber
supporting rails should be covered on top and sides with rubber or PVC at least 1/16" thick.
10. No metal should be in contact with plastic cell containers.
11. Battery connection links should be kept as short as practical, terminals' should be cleaned and the
connecting lugs firmly tightened using stainless steel bolts - do not over tighten.
Grease-impregnated felt washers should be placed under the lugs to arrest corrosion.
The interconnecting lug faces on Telecom type batteries must be cleaned, and if necessary squared with a
coarse file.
The lugs are bolted together, the lug, bolt and nut being lightly coated with petroleum jelly before
assembly. The correct size spanner must be used; pliers or grips must not be used or damage may result.
Nuts must not be over-tightened.
12. During normal battery life, positive plates may expand and increase in length by 5%. Intercell
connections must therefore be soft lead or flexible. Heavy bus bars or charging leads must be able to
accommodate some movement.

MONITORING AND MAINTENANCE


A battery bank will need to be monitored and will need a certain amount of attention from time to time.
First of all, a battery bank most be charged and remain as fully charged as possible. It is advisable in a
home power situation that you have an amp-meter to show the rate of charge and a volt- meter to give
some idea of the state of the batteries. Both of these meters should be mounted in such a position that they
noticed frequently. This strategy will make you more familiar with what to expect and make you aware of
any problem when it arises, such as no amps showing when the batteries are supposed to be charging.
You will need to- take note of the rate of water loss of the battery bank and make sure it is topped up
before the level drops to less than one centimeter above plates or to the lower level marked on some
batteries. Bring it to the bottom of the filler wells or to upper (high) level specified by the manufacturers.
Only top the battery up with distilled water or clean rain water collected in plastic or glass. Do not overfill. It is advisable to take specific gravity measurements of all the cells of the battery bank once in a
while with a hydrometer.

CHARGING THE BATTERY


The word "gas" here refers to a gas given off by the acid due to electrolysis of the water. If continued at a
high rate this gassing can be quite a violent boiling action and will result in loss of water and plate
damage.
These points must be born in mind:
1. If a battery is left in a partially discharged state for an extended period, sulphation of the plates will
occur, which if allowed to proceed, results in irreversible loss of capacity.
2. If a cell is maintained at a constant voltage without any cycling, "stratification" of the electrolyte interlayers of differing densities will occur. This can be minimized by occasionally charging the battery to a
gassing voltage (i.e. some bubbling occurs).

CHARGE LEVEL OF BATTERY


The charged or discharged condition of a lead-acid battery is indicated by the colour of the positive plate,
the voltage, and the strength (specific gravity of the electrolyte). A lead-acid battery will self discharge
over a period of time if left standing and not connected to any charging source.

Connecting Batteries in Parallel


How to correctly interconnect multiple batteries to form one larger bank.
Here is a diagram showing the old way of interconnecting 4 batteries to form one larger bank. This is a
method that we still see in many installations.

Notice that the connections to the main installation are all taken from one end, i.e. from the end battery.
The interconnecting leads will have some resistance. It will be low, but it still exists, and at the level of
charge and discharge currents we see in these installations, the resistance will be significant in that it will
have a measurable effect.
Typically the batteries are linked together with 35mm cable in a good installation (often much smaller in
a poor installation). 35mm copper cable has a resistance of around 0.0006 Ohms per metre so the 20cm
length between each battery will have a resistance of 0.00012 Ohms. This, admittedly, is close to nothing.
But add onto this the 0.0002 Ohms for each connection interface (i.e. cable to crimp, crimp to battery post
etc) and we find that the resistance between each battery post is around 0.0015 Ohms.
If we draw 100 amps from this battery bank we will effectively be drawing 25 amps from each battery.
Or so we think.
In actual fact what we find is that more current is drawn from the bottom battery, with the current draw
getting progressively less as we get towards the top of the diagram.
The effect is greater than would be expected.
Whilst this diagram looks simple, the calculation is incredibly difficult to do completely because the
internal resistance of the batteries affects the outcome so much.
However look at where the load would be connected. It is clear that the power coming from the bottom
battery only has to travel through the main connection leads. The power from the next battery up has to
travel through the same main connection leads but in addition also has to travel through the 2
interconnecting leads to the next battery. The next battery up has to go through 4 sets of interconnecting
leads. The top one has to go through 6 sets of interconnecting leads. So the top battery will be providing
much less current than the bottom battery.
During charging exactly the same thing happens, the bottom battery gets charged with a higher current
than the top battery.
The result is that the bottom battery is worked harder, discharged harder, charged harder. It fails earlier.
The batteries are not being treated equally.
Now in all fairness, many people say "but the difference is negligible, the resistances are so small, so the
effect will also be small".
The problem is that in very low resistance circuits (as we have here) huge differences in current can be
produced by tiny variations in battery voltage. I'm not going to produce the calculations here because
they really are quite horrific. I actually used a PC based simulator to produce these results because it is
simply too time consuming to do them by hand.
Battery internal resistance = 0.02 Ohms
Interconnecting lead resistance = 0.0015 Ohms per link
Total load on batteries = 100 amps

The bottom battery provides 35.9 amps of this.


The next battery up provides 26.2 amps.
The next battery up provides 20.4 amps.
The top battery provides 17.8 amps.
So the bottom battery provides over twice the current of the top battery.
This is an enormous imbalance between the batteries. The bottom battery is being worked over twice as
hard as the top battery. The effects of this are rather complex and do not mean that the life of the bottom
battery will be half that of the top battery, because as the bottom battery loses capacity quicker (due to it
being worked harder) the other three batteries will start to take more of the load. But the nett effect is that
the battery bank, as a whole, ages much quicker than with proper balancing.
I have to be honest now and say that when I first did this calculation in about 1990 I completely refused
to believe the results. The results seemed so exaggerated. So much so that I wired up a battery bank and
did the experiment for real, taking real measurements. The calculations were indeed correct.

All that has changed in this diagram is that the main feeds to the rest of the installation are now taken
from diagonally opposite posts.
It is simple to achieve but the difference in the results are truly astounding for such a simple modification.
The connecting leads, in fact, everything else in the installation remains identical.
Also, it doesn't matter which lead (positive or negative) is moved, Whichever is easiest is the correct one
to move.
The results of this modification, when compared to the original diagram are shown below. Only that one
single connection has been moved.
After this simple modification, with the same 100 amp load....
The bottom battery provides 26.7 amps of this.
The next battery up provides 23.2 amps.
The next battery up provides 23.2 amps.
The top battery provides 26.7 amps.
This is quite clearly a massive improvement over the first method. The batteries are much closer to being
correctly balanced. However they are still not perfectly balanced.
How far is it necessary to go to get the matching equal?
Well, the better the quality of the batteries, the more important it becomes. The lower the internal
resistance of the batteries, the more important it is to get them properly balanced.
So that now leaves the question of whether or not there is a wiring method to perfectly balance the
batteries.
Before getting to that, it should be pointed out that doing the calculation is not actually required in order
to arrive at the ultimate interconnection method. I simply did them to show the magnitude of the problem.
In order to get a better balancing it is simply necessary to get the number of interconnecting links as close
as equal between each battery and the final loads.
In the first example the power from the bottom battery passed through no interconnecting links. The top
battery passed through 6 links.

In the 2nd example (the much improved one), the power from the top and bottom battery both passed
through a total of 3 links. That from the middle 2 batteries also both passed through 3 links which begs
the question "why were they not therefore perfectly balanced?". The answer is that some of the links have
to pass more total current and this therefore increases the voltage drop along their length.
And finally we get to the correctly wired version where all the batteries are perfectly balanced.

This looks more complicated.


It is actually quite simple to achieve but requires two extra interconnectng links and two terminal posts.
Note that it is important that all 4 links on each side are the same length otherwise one of the main
benefits (that of equal resistance between each battery and the loads) is lost.
The difference in results between this and the 2nd example are much smaller than the differences between
the 1st and 2nd (which are enormous) but with expensive batteries it might be worth the additional work.
Most people (myself included) don't consider the expense and time to be worthwhile unless expensive
batteries are being fitted or if the number of batteries exceeds 8.
However there really is no excuse whatsoever (apart from, perhaps, incompetence or laziness) for using
the first example given at the top of this page.
This final method achieves perfect balancing between all four (or indeed any other number of) batteries.
I think I am right in saying that this is the only example I have ever come across where doing something
the correct way actually looks less elegant than doing it incorrectly.
Finally, if you only have 2 batteries, then simply linking them together and taking the main feeds from
diagonally opposite corners cannot be improved upon.
Once the number of batteries gets to 3 or more then these other methods have to be looked at.
With a large number of batteries it may be necessary to go to the final method shown above.
Even with 8 batteries it is possible to get reasonable balancing by placing the main "take off" feeds from
somewhere down the chain instead of from the end batteries. Remember, count the number of links each
battery needs to run through to reach the final loads and get these as equal as possible.
Finally, if your battery bank has various take off points on different batteries, change it now! It is
extremely bad practice. Not only does it mess up the battery balancing, it also makes trouble shooting
very much more complicated and looks awful.

Cell Balancing
In multicell batteries, because of the larger number of cells used, we can expect that they will be subject
to a higher failure rate than single cell batteries. The more cells used, the greater the opportunities to fail
and the worse the reliability.
The problems can be compounded if parallel packs of cells are required to achieve the desired capacity or
power levels. With a battery made up from n cells, the failure rate for the battery will be n times the
failure rate of the individual cells.

All cells are not created equal


The potential failure rate is even worse than this however due to the possibility of interactions between
the cells. Because of production tolerances, uneven temperature distribution and differences in the ageing

characteristics of particular cells it is possible that individual cells in a series chain could become
overstressed leading to premature failure of the cell.
During the charging cycle, if there is a degraded cell in the chain with a diminished capacity, there is a
danger that once it has reached its full charge it will be subject to overcharging until the rest of the cells in
the chain reach their full charge. The result is temperature and pressure build up and possible damage to
the cell. During discharging, it is even possible for the voltage on the weaker cells to be reversed as they
become fully discharged before the rest of the cells resulting in failure of the cell.
Balancing is less of a problem with parallel chains which tend to be self balancing.
Because Lead acid and NiMH cells can withstand a level of over-voltage without sustaining permanent
damage, a degree of cell balancing or charge equalisation can occur naturally with these technologies
simply by prolonging the charging time since the fully charged cells will release energy by gassing until
the weaker cells reach their full charge. Unlike Lithium cells which can not tolerate over-voltages.
Although the problem is reduced with Lead acid NiMH batteries and some other cell chemistries, it is not
completely eliminated and solutions must be found for most multicell applications.
No matter what battery management techniques are used, the failure rate or cycle life of a multicell
battery will always be worse than the quoted failure rate or cycle life of the single cells used to make up
the battery.
Once a cell has failed, the entire battery must be replaced and the consequences are extremely costly.
Replacing individual failed cells does not solve the problem since the characteristics of a fresh cell would
be quite different from the aged cells in the chain and failure would soon occur once more. Some degree
of refurbishment is possible by cannibalising batteries of similar age and usage but it can never achieve
the level of cell matching and reliability possible with new cells.

Cell selection
The first approach to solving this problem should be to avoid it if possible through cell selection.
Batteries should be constructed from matched cells, preferably from the same manufacturing batch.
Testing can be employed to classify and select cells into groups with tighter tolerance spreads to
minimise variability within groups.

Pack construction
Another important avoidance action is to ensure at all times an even temperature distribution across all
cells in the battery. On the other hand, if the cells are concentrated in one large block, the outer cells in
contact with ambient air may run cooler than the inner cells which are surrounded by warmer cells unless
steps are taken to provide an air (or other coolant) flow to remove heat from the hotter cells.

Cell equalisation
To provide a dynamic solution to this problem which takes into account the ageing and operating
conditions of the cells, the BMS may incorporate a Cell Balancing scheme to prevent individual cells
from becoming overstressed. These systems monitor the State of Charge (SOC) of each cell, or for less
critical, low cost applications, simply the voltage across, each cell in the chain. Switching circuits then
control the charge applied to each individual cell in the chain during the charging process to equalise the
charge on all the cells in the pack.
Several Cell Balancing schemes have been proposed and there are trade-offs between the charging times,
efficiency losses and the cost of components.

Active balancing
Active cell balancing methods remove charge from one or more high cells and deliver the charge to one
or more low cells. Since it is impractical to provide independent charging for all the individual cells
simultaneously, the balancing charge must be applied sequentially. Taking into account the charging
times for each cell, the equalisation process is also very time consuming with charging times measured in
hours. Some active cell balancing schemes are designed to halt the charging of the fully charged cells and
continue charging the weaker cells till they reach full charge thus maximising the battery's charge
capacity.

Passive balancing
Dissipative techniques find the high cells in the pack, and remove excess energy through a resistive
element until their charges match the low cells. Some passive balancing schemes stop charging altogether
when the first cell is fully charged, then discharge the fully charged cells into a load until they reach the

same charge level as the weaker cells. Other schemes are designed continue charging till all the cells are
fully charged but to limit the voltage which can be applied to individual cells and to bypass the cells when
this voltage has been reached.

Charge limiting
A crude way of protecting the battery from the effects of cell imbalances is to simply switch off the
charger when the first cell reaches the voltage which represents its fully charged state and to disconnect
the battery when the lowest cell voltage reaches its cut off point during discharging.
This will unfortunately terminate the charging before all of the cells have reached their full charge or cut
off the power prematurely during discharge leaving unused capacity in the good cells. It thus reduces the
effective capacity of the battery. Without the benefits of cell balancing, cycle life could also be reduced,
however for well matched cells operating in an even temperature environment, the effect of these
compromises could be acceptable.
All of these balancing techniques depend on being able to determine the state of charge of the individual
cells in the chain. Several methods for determining the state of charge are described on the SOC page.
The simplest of these methods uses the cell voltage as an indication of the state of charge. The main
advantage of this method is that it prevents overcharging of individual cells, however it can be prone to
error. A cell may reach its cut off voltage before the others in the chain, not because it is fully charged but
because its internal impedance is higher than the other cells. In this case the cell will actually have a
lower charge than the other cells. It will thus be subject to greater stress during discharge and repeated
cycling will eventually provoke failure of the cell.
More precise methods use process called Coulomb counting and take account of the temperature and age
of the cell as well as the cell voltage.
For batteries with less than 10 cells, where low initial cost is the main objective, or where the cost of
replacing a failed battery is not considered prohibitive, cell balancing is sometimes dispensed with
altogether and long cycle life is achieved by restricting the permitted DOD. This avoids the cost and
complexity of the cell balancing electronics but the trade off is inefficient use of cell capacity.
Whether or not the battery employs cell balancing, it should always incorporate fail safe cell protection
circuits.

Lead-Acid Battery State of Charge vs.Voltage


A battery voltmeter is the most basic system instrument. Battery voltmeters are inexpensive, easy to
install, and can provide a wealth of system information to renewable energy users, or anyone who
depends on a battery.
Why a Voltmeter?
Ten years ago, voltmeters were all we had for information about our systems. Ampere-hour meters that
calculated battery efficiency were a pipe dream. Even now, small systems cannot justify the additional
expense and complexity of the new sophisticated battery state of charge (SOC) instruments. The
voltmeter is always there, consumes virtually no power, and tells me at a glance whats happening with
our system.
Reading a battery voltmeter and turning that information into a reliable assessment of the batterys state
of charge is like tracking an animal by its footprints. Tracking requires noticing small details and
extrapolating information from these details. A tracker uses his knowledge of the animals habits. A
tracker considers the weather and season. A trackers knowledge of his subject and its environment
allows him to predict the actions ofhis subject.
After watching the voltmeter for a few of the batterys charge/discharge cycles, the user gets a idea of his
batterys voltage profiles. After watching the voltmeter for a season or two, the user learns how to relate
the effects of temperature and current on his batterys voltage. Just like the behavior of animals vary with
type and location, the behavior of batteries differ with type and operating environment.
What Kind of Voltmeter?

It really doesnt matter what type of voltmeter you use to measure your batterys voltage. Better
instruments yield more accurate measurements with higher resolution. Differences in battery voltage of
0.1 VDC are significant, so the instrument should have a basic accuracy at least 0.5% or better.
Accurate analog battery voltmeters can be purchased for under 20. Digital multimeters cost from 5 to
100 and perform highly accurate voltage measurements and much more besides. Or you can homebrew
an expanded scale analog battery voltmeter. You can homebrew an LED battery meter. Any of these
instruments will give you the voltage measurement you need.
Installation of a battery voltmeter is easy. Just connect it to the batterys main positive and negative buss
or terminals. Be sure to get the polarity right because analog meters can be damaged by reverse polarity.
Since the battery voltmeter consumes very little power, the wires feeding it can be small (18 gauge
copper or smaller).
Reading the Curves
The data presented here on the graphs was generated from our set of Trojan L-16W deep cycle lead-acid
batteries. Each Trojan L-16 battery is composed of three series connected, 350 Ampere-hour, lead-acid
cells. The graphs and the data here relates to six of these lead-acid cells in series forming a 12 Volt
battery. Those of you using a 24 Volt system with twelve leadacid cells in series must multiply the
voltage in the text and on the charts by two.
The voltage versus state of charge (SOC) profiles will match those of similarly constructed cells. Other
types of lead acid cells, like car batteries, lead-calcium cells, and RV deep cycle batteries will have
different charge/discharge curves. I offer these graphs as examples of what to look for with your battery.
While specific voltage vs. SOC points will vary from battery type to battery type, the shape and
relationship of the curves is similar for all deep cycle lead-acid technologies.
Current and Batteries and Ohms Law
Battery voltage can be affected by three factors state of charge, current, and temperature. State of
charge is what we are trying to find out, so that leaves current and temperature as factors to reckon with.
Current means the rate of electron flow through the battery caused by either charge or discharge. Every
electrochemical cell has internal resistance. As current moves through the cell, the cells voltage changes
because of this internal cell resistance. When the cell is being recharged, current flow causes the cells
voltage to rise. The higher the recharging current the higher the voltage rise. As the cell is discharged, the
discharging current causes the cells voltage to drop.
The higher the discharging current, the greater the batterys battery depression. This holds true for all
electrochemical cells regardless of type, size, or environment. While absolute values vary widely between
different acid and alkaline technologies, the relationship between current flow and cell voltage remains
constant.
The graphs show a variety of recharge and discharge rates from C/5 to C/100. This C/XX number is
actually a rate of charge or discharge in Amperes proportioned to the capacity of the battery. For
example, consider a battery of 100 Ampere-hours. If you divide this Amperehour capacity by 10 hours,
then you get a charge (or discharge) rate of 10 Amperes.
Ten Amperes is a C/10 charge (or discharge) rate for a 100 Ampere-hour battery. Consider another
battery of 500 Ampere-hours capacity. Here a C/10 rate would be 50 Amperes. While the absolute values
of the charge (or discharge) currents is different between the two batteries of different capacity, their
effect on the batterys voltage is the same. The currents are in the same proportion to the batteries
capacity.

If voltage is to be related to battery state of charge, then we must compensate for voltage variation due to
current movement through the battery. Hence there are a variety of curves on both the charge and
discharge graphs.
Included on the charge graph is a gray curve entitled Rest. This rest curve is a generic representation of
six electrochemical cell has internal resistance. As current moves through the cell, the cells voltage
changes because of this internal cell resistance. When the cell is being recharged, current flow causes the
cells voltage to rise. The higher the recharging current the higher the voltage rise.
As the cell is discharged, the discharging current causes the cells voltage to drop. The higher the
discharging current, the greater the batterys battery depression. This holds true for all electrochemical
cells regardless of type, size, or environment. While absolute values vary widely between different acid
and alkaline technologies, the relationship between current flow and cell voltage remains constant.
The graphs show a variety of recharge and discharge rates from C/5 to C/100. This C/XX number is
actually a rate of charge or discharge in Amperes proportioned to the capacity of the battery. For
example, consider a battery of 100 Ampere-hours. If you divide this Amperepronounced enough to
distinctly change not only the battery voltage vs. SOC profile, but also its useful Ampere-hour capacity.
The discharge voltage curves may be depressed by as much as 0.5 VDC from those shown on the graph.
Charge voltages will be elevated by as much as 0.5 VDC for a cold 12 Volt lead-acid battery.

Lead-acid Internal Resistance and SOC


In lead-acid cells, the electrolyte (sulfuric acid) participates in the cells normal charge/discharge
reactions. As the cells are discharged, the sulfate ions are bonded to the plates sulfuric acid leaves the
electrolyte. The process is reversed when the cell is recharged.
A fully charged lead-acid cell has an electrolyte that is a 25% solution of sulfuric acid in water (specific
gravity about 1.26). A fully discharged lead-acid cell has lead-acid cells in series and at Rest. At Rest
means that no current is moving through the cells, i.e., that they are neither being charged or discharged.
Determining a batterys state of charge from voltage measurement is vague enough if current is moving
through the battery. The vagaries increase exponentially if no current is moving through the battery. This
is why this curve is gray.
Temperature and Batteries
The lead acid reaction is temperature sensitive. Cooling the cell changes its voltage vs. SOC profile. As
the lead-acid battery cools, its internal resistance increases. This means that voltage elevation under
recharging is increased in cold cells. The same internal resistance increase produces increased voltage
depression in cold cells when discharged.
At 32F (0C), the effect of temperature becomes virtually no sulfuric acid in its almost pure water
electrolyte (specific gravity about 1.00). As the sulfuric acid concentration in the electrolyte changes so
does the electrical resistance of the electrolyte, which in turn changes the internal resistance of the entire
cell.
The bottom line is that the internal resistance of all lead-acid cells changes with the cell's state of charge.
This characteristic gives the lead-acid reaction its particular shape or signature on the voltage vs. SOC
graphs. This signature is unique - very different from alkaline cells whose electrolyte resistance remains
constant regardless of SOC. The shape of the lead-acid curves makes it possible to use a voltmeter to
determine a battery's state of charge.
Reading the Tracks
The more you understand the relationship between battery voltage and real life events like current
movement and temperature, the more information transferred by a simple voltage measurement. Your
battery savvy here is worth more than a 10 voltmeter.

Battery FAQ
Q: What is battery sulfation and why does it cause battery failure ?
A: Each time a battery is discharged some of the sulfuric acid in the electrolyte is converted to lead
sulfate crystals which form on the plates. These crystals interfere with the chemical reaction neccessary
for proper battery operation. When the battery is recharged MOST (but not all) the sulfates are
reconverted back into sulfuric acid. This is why a discharged battery will have a low Specific Gravity
(SG) reading. Over time the sulfates which are not reconstituted as acid form stubborn patches of large
crystal growths that will grow and eventually choke the life out of a lead acid battery. These crystal
growths if not stopped will act like ice expanding the cases (bulges), breaking the plates and plate and cell
interconnects and causing shorts.
Q: Why does a battery die when it sits ?
A: All lead acid batteries will self discharge forming sulfates in the process. The degree of self discharge
will depend on temperature, plate chemistry, and how clean the battery is and how it is stored.
Q: Is there any truth to storing a battery on the ground being a bad idea ?
A: Yes, When stored on the ground the battery will assume the same temperature as the ground. When
warmer moist air comes in contact with the battery case, moisture can condense out (like an iced drink in
the summer) and increase the amount of self discharge across the top of the case (between the posts). To
see this in action take a voltmeter and attach one lead to a post and drag the other lead across the top. You
will note an increasing voltage as the probe heads to the other post. To reduce this aspect of self discharge
always store batteries on a shelf or on some form of insulation.

Lead-Acid Battery De-sulfator

SCHEMATIC DIAGRAM and OPERATION


This circuit shown at below is essentially a switching DC-to-DC converter that steps a DC voltage up to a
higher level. It takes power from the battery and pulses it back into the battery. The pulse rate is set by the
555 timer chip,U1 which switches the MOSFET at a 1 kHz rate. When Q1 is in the non-conducting state,
current is drawn from the battery through L2 so it charges capacitor C4 slowly. Q1 is then switched on
for 50 microseconds, causing C4 to discharge through L1. When Q1 is switched off again, the stored
inductive energy in L1 pulses back into the battery through diode D1.
This pulse of current can be as high as 6 amps. The use of an inductor to supply this high voltage pulse is
what makes it possible to restore a badly sulfated battery with a high internal resistance. The peak voltage
applied across the battery can be as high as 50 volts. This voltage will decrease as the batterys internal
resistance declines or its function is restored to normal.

NOTES
There is no reverse polarity protection for this circuit so label the leads well. Connect them backwards
and you will blow components. Connect positive of the pulse charger to positive of the battery!
There is an issue with electrical equipment connected across the battery; lights, radio, etc. Depending on
the impedance of this equipment, it may absorb some of the pulse energy, thereby minimizing the effect
on the battery. To solve this problem, slip a ferrite toroid core over each positive battery lead (right at the
battery) going to the other equipment.
Exclude the pulse charger lead. The ferrite core will increase the high frequency resistance without
affecting the DC performance of the circuitry. Hence all the pulse energy will enter the battery and not be
consumed by the electrical loads connected to it.

Keep the lead length from the pulse charger to the battery as short as possible to minimize RF radiation
and power loss in the wires. A faint audio tone might be heard when the circuit is in operation.
Pulse energy happens at less than 100% efficiency. The circuit draws about 40 MA from the battery so
some additional charging must be applied to offset this. A solar panel would be perfect.

Parts List:
Q1 IRF9540 P channel MOSFET
IC1 LM555CN Timer IC
D1 FR607 Fast recovery diode, >6 A, 100 V
C1 30F, 16 V Electrolytic
C2 0.0022F Disk ceramic
C3 0.047F Disk ceramic
C4 100F, 16 V Electrolytic, low impedance type
R1 470 k. 1/4 W
R2 22 k. 1/4 W
R3 330 . 1/4 W
R4 330 . 1/4 W
L1 220H (nominal) Ferrite inductor, 6+ A peak
L2 1000H Ferrite choke, 100 mA
Case Aluminum project box
Clip leads Alligator type, insulated (RS)
Board material 0.1" spaced copper pads

Battery Distribution Systems


All battery storage systems should be set up and used safely. All connections should be via fuses or
circuit breakers and all cabling rated for it's purpose and fuse rating. Good quality battery isolation
switches are recommended and suitable voltage and current meters (ammeters) should be installed.
In Figure 1, the various charging sources are fed via ammeters to show charging current and suitable
fuses to the battery charging regulator (marked 'voltage regulator'). The output of this is fed via switches
(or diodes) to the battery banks.

Figure 1. A typical small/medium size installation for cottage/boat/truck/caravan etc.


The output is fed via switching (and preferably overload cut-outs, not shown) to the various outputs.
Battery outputs must always be fed through low voltage cut-outs where appropriate (most modern
inverters incorporate this function internally), and via appropriately rated fuses or circuit breakers.
It is also recommend that each battery has it's own high current breaker or fuse mounted close to its
terminals.
High current switches for two battery banks can be obtained from boat chandlers, as can many other
useful items.
These outputs may be sent via different ammeters (small and large), or even from separate battery banks,
with low power 'everyday' equipment on one battery and large inverters and power circuits on another for
example.
Along with the ammeters there is also shown, a switchable voltmeter to check the state of the batteries.

Note that it is better to have two or more smaller banks of batteries that one large one. Batteries like to be
discharged and charged, so one can be drained down toward its safe end point and then swapped for the
other charged battery. The charging source is then applied to that battery until it is charged.
A back-up battery is also shown and this could be one that is designed specifically for standby use and
will generally be smaller than the main batteries.
It is worth noting that if you intend to run mains voltage equipment via an inverter then it must be
double insulated (i.e. the mains lead does not contain an earth), if the equipment requires an earth then
this should be provided, if necessary using an earth spike. This is a copper covered steel pole specially
designed for the job and is available from all electrical wholesalers. (They do last a long time even when
banged into the average field etc., and removed umpteen times!).

If used within a metal bodied vehicle, then the bodywork can be used, (the neutral and earth of an inverter
are usually bonded together and are often bonded as well to the battery negative, check yours with a
multimeter). NOTE: This is only safe whilst used within the confines of the vehicle, if the equipment,
any part of, or anything connected that may possibly be conductive, is accessable from outside the
vehicle, it must be earthed properly

12V lead acid batteries - Voltage references:


Lead Acid (liquid electrolyte): Charge/discharge voltages:
SOC= State Of Charge
11.7V - Prewarning: SOC<40%
11.1V - Disconnection: SOC<30%
12.4V - Reconnection: SOC>50%
11.7V - Commence Equalisation charge (14.7V) (this means that if the battery has reached this level of
discharge or beyond, it should ideally receive a charge at 14.7V)
12.6V - SOC<70% Cycle charge (14.4V) (this means that if the battery has been discharged beyond this
point, it should ideally receive a charge at 14.4V, the standard charge voltage like that on a car
alternator).
13.7V - Final charge voltage
Temperature compensation: -4mV/K/Cell (ideally, but not often used!)

You might also like