You are on page 1of 170

Energy and the New Reality, Volume 2:

C-Free Energy Supply

Chapter 3: Wind Energy

L. D. Danny Harvey
harvey@geog.utoronto.ca

Publisher: Earthscan, UK
Homepage: www.earthscan.co.uk/?tabid=101808

This material is intended for use in lectures, presentations and as


handouts to students, and is provided in Powerpoint format so as to
allow customization for the individual needs of course instructors.
Permission of the author and publisher is required for any other usage.
Please see www.earthscan.co.uk for contact details.
Figure 3.1b Growth in total wind energy capacity, 2000-
2016

By comparison, the total global electricity generation capacity is about 5500 GW


Figure 3.1a Annual additions to wind energy capacity, 2000-2016

By comparison, the total existing nuclear + hydro + natural gas electricity generating capacity in
Ontario is just under 30 GW (with another 10 GW of wind capacity and 2 GW of solar)
Note that about half of all the wind turbines installed in the
world in 2016 were installed in China. However, China seems
to be building up wind energy too fast – there are major
problems getting the turbines connected to the grid or being
able to transmit the generated electricity. As a result, wind
farm output frequently has to be “curtailed”. Curtailment rates
are 15-20% in various Chinese provinces, compared to about
2% in the US. Some wind farms are not yet connected to the
grid. Between unconnected wind farms and curtailed output,
wind electricity production in China in 2012 was only about
65% that expected (Lam et al., 2016, ERL).

As well, many of the Chinese wind manufacturers are state-


owned and have been willing to sell wind turbines at a loss in
order to gain market share or to meet government targets.
Figure 3.3 Wind farm at Pincher creek, Alberta

Source: Garry Sowerby


Ta b le 3.1
Table 1 0 .1Market
M a rkshares
e t s h a re
of sthe
o f world’s
th e w o leading
rld ’s le ad in g
wind
wturbine
in d tu rb in e m a n u fa c tuSource:
manufacturers. re rs . S oBTM
u rc eConsult
: B T M Press
CRelease,
o n su lt PMarch
re ss R2007
e le a s e, M a rc h 2 0 0 7 .
M a rk et
Ve n d o r S h are in C o u n try
2006

Ve sta s 2 6 .5 D e n m a rk
G a m e sa 1 4 .6 S p ain
G E W in d 1 4 .6 U SA
E n e rc o n 1 4 .5 G e rm a n y
S u zlo n 7 .2 In d ia
S ie m a n s 6 .9 G e rm a n y
N o rd e x 3 .2 G e rm a n y
R epow er 3 .0 G e rm a n y
A c c o n ia 2 .6 S p ain
G o ld w in d 2 .6 C h in a
O th e rs 4 .3
Market share of global wind turbine sales in 2016

Source: Renewables 2017, Global Status Report

Global employment in the wind energy sector: 1.1 million (REW Jan-Feb 2017, p 20)
Components of a Wind Turbine

• Foundation
• Tower
• Rotor
• Nacelle
• Gearbox (in older units)
• High speed shaft
• Generator
• Control system, cooling
unit, anemometer
• Yaw mechanism
Turbine characteristics

• Rotor diameter – up to 120 m


• Hub height – up to 120 m
• Peak electrical power output – up to 6 MW now,
up to 15 MW foreseen (offshore)
• Cut-in wind speed (typically 3-4 m/s)
• Rated wind speed (typically 15 m/s)
• Cut-out wind speed (typically 25 m/s)
Turbine size (power capacity)

• Onshore – typically 2 MW, some up to 4 MW, but


restricted due to the logistical difficulties of transporting
blades by road
• Offshore – many 6-7 MW turbines are available, the
largest currently available is the 9.5 MW Vestas V-164,
and 15 MW turbines are currently envisaged
• Building larger turbines for offshore is regarded as one of
the keys to reducing the still high unit cost of offshore
wind energy
Figure 3.4 Progression of rotor sizes over time
Source: Renewable Energy World, Jan-Feb 2017
Note: falling specific power means that rotor area is increasing faster than generator capacity,
but this makes it possible to generate more electricity at a given wind speed and to reach full
power at a lower wind speed.

Source: Renewable Energy World, Jan-Feb 2017


The preceding slide shows wind turbines for
offshore applications growing in maximum
size from 4.1 MW in 2015 to 11 MW by
2030, but there have been discussions of
offshore turbines reaching 15 MW and even
50 MW capacity! Bigger turbines require
development of ultra-light components
(using C fibres, for example)
Figure 3.7a Power curves for wind turbines with 80-m,
87-m, and 90-m rotors and a 2.0-MW generator
2000

1500
Power Output (kW)

Gamesa G80-2.0
1000
Gamesa G87-2.0

Gamesa G90-2.0

500

0
0 5 10 15 20 25
Wind Speed (m/s)
Wind turbine aerodynamics

• Lift, not a pushing force, is what makes the rotor


rotate
• Thus, the aerodynamics of a wind turbine have
much in common with the aerodynamics of an
airplane wing
Figure 3.8 Airflow Past Wing

Faster air flow - less pressure on wing

Net Force

Slower air flow - greater pressure on wing


Figure 3.9 Forces acting on a turbine rotor blade

D
L

D co s

V W

90
tip radius R

Source: Danish Wind Turbine Manufacturers’ Association


Efficiency of a wind turbine: this is the ratio of the
electrical power produced (W) to the power of the
wind passing through the area swept by the rotor
blades. It is the product of three factors:

• Aerodynamic efficiency (ratio of mechanical power of the


rotor to wind power)
• Mechanical efficiency (ratio of mechanical power of the
generator axis to the mechanical power of the rotor axis)
• Electrical efficiency (ratio of electrical power fed into the
grid to the mechanical power of the generator axis)
• The maximum possible aerodynamic efficiency,
as given by Betz’ Law, is 59.3%, and occurs if
the turbine slows the wind down to 2/3 of its
original speed. The aerodynamic efficiency of a
real turbine varies with wind speed, having a
typical peak value of 44% and a typical value
averaged over all wind speeds of 25%
• A typical mechanical efficiency is 96-99%
• A typical electrical efficiency is 96-97%
• Multiply the efficiencies (expressed as a fraction)
to get the overall efficiency
Figure 3.10: Variation of power output and efficiency
with wind speed for the Nordex N90-2.3 turbine
2500 0.5

2000 0.4

1500 0.3
Output (kW)

Efficiency
1000 0.2

500 0.1

0 0.0
1 3 5 7 9 11 13 15 17 19 21 23 25
Wind Speed (m/s)
Turbine generators

Electricity (the flow of electrons) is produced by a


changing magnetic field. An electrical generator
consists of 3 stationary magnets as part of a
stator, which is a circular ring within which a rotor
with further magnets rotates. Wire is wound
around each stator magnet. The rotating rotor
magnet induces fluctuating voltage in each of the
stator windings, and since each stator magnet is
offset by 1/3 of the circumference, the current
produced in each wire is offset by 1/3 of a cycle
from the adjacent wire. The produces what is
called 3-phase AC.
Recap Volume 1 Figure 3.2
Two Pole Synchronous Generator
Three Phase AC Current

1.0
Voltage

0.0

-1.0
0 5 10 15 20
Time (ms)
The frequency of the electricity (cycles per second)
produced by the generator rotor with one magnet would
equal the rotor frequency, that from a rotor with 2 magnets
would be twice the rotor frequency (since the rotor would
need to rotate only half a circle to go from one north pole to
the next), and so on. Electricity in North America has a
frequency of 60 Hz (60 cycles per second), but the wind
turbine rotor might rotate at only 10 cycles per minute – so
(in most wind turbines), a system of gears is needed to
step up from the turbine rotor speed to the required
generator rotor speed, and then electronics is used to get
an exact match in both the frequency and phase between
the turbine electricity output and those of the grid to which it
is connected.
There are two basic variants of the generator:

• An asynchronous or induction generator, in which the


magnets in the stator are induced by an electrical current
supplied from the grid to which the turbine is connected,
and
• A synchronous generator, which may or may not have
permanent magnets (not requiring any electrical current for
magnetization) in the stator
• Asynchronous generators can be further divided into the squirrel
cage induction generator (SCIG) and the wound rotor induction
generator (WRIG), which differ in how the rotor is constructed.
They are illustrated in the next slide.
Schematic diagram of a squirrel cage
induction generator (SCIG)
Squirrel cage and wound rotor induction generators
The choice of generator strongly affects how the wind
turbine rotor rotation rate varies with wind speed, which in
turns affects wear and tear, efficiency, and noise
Asynchronous SCIGs or WRIGs
• If the generator rotor were to rotate at the same frequency
as the electric field in the stator, no electricity would be
produced
• When the rotor of the generator rotates faster than the
stator, a strong current is induced
• The harder one cranks on the rotor, the more power that is
transferred as electromagnetic force to the stator, converted
to electricity, and fed to the grid
• The difference between the rotor and the stator frequencies
is called the slip,
slip and at peak power it is only a few % (i.e.,
the rotation rate of the turbine rotor varies by only a few %
between zero and maximum power output)
To allow some variation in rotor speed (from -30% to
+40% from the synchronized speed), the doubly-fed
induction generator (DFIG) was developed.
• “Doubly-fed” means both the stator and the rotor (of the WRIG type)
are connected to the grid, the stator as usual with a current that
induces magnetization
• The rotor directly feds about 30% of the power output through a
partial-scale frequency convertor that produces AC current such that
when it is blended with the stator output, the resulting current exactly
matches the frequency of the grid even as the rotor rotation rate
varies.
Schematic diagram of a
doubly-fed induction generator (DFIG)

Source: Wikipedia
The latest step in the evolution of turbine
generators is the permanent magnet
synchronous generator. It is

• More expensive and mechanically more complicated


than an induction generator
• The permanent magnet requires (at present) rare earth
elements neodymium (Nd) and dysprosium (Dy) that
could be in short supply in the future
• However, such turbines are more efficient, allow a wide
variation in turbine rotor rotation rate, and lend
themselves to elimination of the gear box altogether
• Presently, about 10% (by capacity) of wind turbines
being sold use permanent magnets, but this fraction will
likely grow
Aside: the production of Nd and Dy for permanent magnet
synchronous wind turbine generators is anything but
“clean”, as illustrated by the toxic waste dump next to a
processing plant in China shown below

Source: Zhang (2013, Peak Nd, MSc thesis)


Note, from below, that the AC output of the wind turbine (of varying frequency)
is converted to DC, then converted back to AC of the exact frequency and
phase needed to match the grid. However, if we were to transmit the electricity
as DC (as discussed later), we could skip the DC-AC conversion at this point.
Characteristics of wind

• Variation of mean wind speed with height


• Variation of turbulence intensity with height
• Weibull probability distribution function for wind
speed
Figure 3.11 Logarithmic velocity profile

70

z
60 U (z) = u * ln z0
K
Z2 U(z 1 )ln(z 2) - U (z 2)ln(z 1)
50
1n (z 0 ) =
U(z 1 ) - U (z 2)
H eig ht, z (m )

40

30 U plots as a
straight line on
20 semi-log paper,
with slope u*/ĸ.
Z1
zo is the height at
10
which U
extrapolates to
0 zero
0 5 10 15
U (z) (m /s)
An alternative mathematical representation of the
variation of wind speed with height is using a
power relationship,

Uh/Uref = (H/href)n

The logarithmic relationship is theoretically valid in


a neutral (neither stable nor unstable) atmosphere
only.

The power relationship has no theoretical basis but


provides a good fit to observed atmospheric wind
profiles
Power output from a wind turbine
• Kinetic energy of a moving mass = ½ mv2
• Power density of wind (rate of flow of energy (watts)
across a plane perpendicular to the wind, per m2 of
plane area) = ½ ρV3
• The efficiency of a wind turbine is defined as the
ratio of power output to the power of the wind in the
area swept by the rotating rotor. Thus,
• Power output of a wind turbine
= efficiency x swept area x power density of wind, or
P=1/2 η(πR2) ρ V3
Weibull Distribution Function
• Gives the probability of a wind speed occurring per
unit of wind-speed interval
• Thus, the units are 1/(m/s)
• The value of the function times the width of the
interval gives the probability of the wind speed
occurring in that interval
• The function is

f(u)=k/c(u/c)k-1exp(-(u/c)k)

where c is the scale parameter and k is the shape


parameter
Figure 3.15 Weibull wind speed distribution with c=5
m/s and k=1.6

0.16
0.14 Mode
Median
0.12
Probability (m/s)-1

0.10 Mean
0.08
0.06
0.04
0.02
0.00
0 5 10 15 20 25
U (m/s)
Comparison Weibull probability function and the
corresponding cumulative probability for c=5 m/s and k=1.6
Note
• The mode is the wind speed where the peak in the
probability curve occurs. It can be read off the
graph, or computed by setting the derivative of the
Weibull distribution function to zero and solving for u
• The median is the wind speed such that wind
speeds greater or lessor than that wind speed occur
50% of the time – so it is the wind speed such that
the cumulative probability = 0.5. It can be found by
setting F(u)=0.5, where F(u) is the cumulative
distribution function, and solving for u.
Figure 3.14 Distribution of best-fit Weibull scale factor
(c) and shape factor (k) deduced from observed wind
velocity variations at various sites
4

3
Shape Factor

0
0 2 4 6 8 10 12
Scale Factor (m/s)
Figure 3.16 Two Weibull wind speed probability
distributions with almost the same mean wind speed –
but very different mean wind power!
Because wind power varies non-
linearly with wind speed
• The mean (average) wind power for a given
mean wind speed depends on the shape of the
probability distribution on either side of the mean
wind speed
• The mean wind power (based on wind power
computed at many different wind speeds and
then weighted by the probabilities) is about twice
the wind power computed once at the mean
wind speed
Figure 3.17 Mean wind power vs mean wind speed.
A smaller k means a more spread out wind speed distribution, so more
winds at both very high and very low wind speeds, but the high wind
speeds disproportionately contribute to wind power (due to the cubic
dependence), so the mean wind power is greater at a given mean wind
speed with smaller k

1500
Mean Wind Power Density (W/m2)

k =1.6

1000 k =2.0

k =2.4

500

0
3 4 5 6 7 8 9 10
Mean Wind Speed (m/s)
Table 3.3. Comparison of wind power computed at the
average wind speed with the average wind power
computed over a distribution of wind speeds giving the
same average wind speed.

Av er a g e W in d p o w e r W in d p o w e r (W /m 2 ) co m p u ted o v e r a d istrib u tio n o f


w in d sp ee d (W /m 2 ) c o m p u te d w in d sp ee d s h a v in g th e sa m e a v er a g e w in d sp eed as
(m /s) a t a v er a ge w in d u se d in c o lu m n 2
sp e ed k = 1 .6 k = 2 .0 k = 2 .4
3 16 40 31 26
5 75 187 144 122
7 207 513 394 336
9 439 1088 838 715
Mean Efficiency

• The power output at any given wind speed is


given by the wind power x swept area x
efficiency, so the efficiencies matter more when
the wind power is larger than when it is smaller
• Thus, the appropriate mean efficiency involves
the efficiency at each wind speed times the
probability of that wind speed interval times the
wind power at that wind speed, divided by the
mean wind power
Figure 3.18a Mean efficiency vs wind speed, computed from the
turbine power curve and the Weibull wind speed probability
distribution using 3 different shape parameters
0.50

0.40
Mean Efficiency

0.30

k=1.6
0.20
k=2.0
0.10
k=2.4

0.00
3 6 9 12
Mean wind speed (m/s)
Capacity Factor
• This is the mean (average) power output of the
turbine divided by the peak (or rated) power
output
• The mean power output is computed as the
power output in the centre of each wind speed
interval, times the probability of that interval,
summed over all intervals and divided by the
total probability (which is 1.0)
Figure 3.19a Variation of capacity factor with wind
speed for 3 different Weibull shape parameters
0.7

0.6 k=2.0

0.5
Capacity Factor

0.4

0.3
k=1.6
0.2

0.1 k=2.4

0.0
3 6 9 12
Mean wind speed (m/s)
RECALL: Figure 3.7a Power curves for wind turbines with
80-m, 87-m, and 90-m rotors and a 2.0-MW generator
2000

1500
Power Output (kW)

Gamesa G80-2.0
1000
Gamesa G87-2.0

Gamesa G90-2.0

500

0
0 5 10 15 20 25
Wind Speed (m/s)
Figure 3.19b Variation of capacity factor with wind
speed for three different turbines
0.7

0.6
N100-2.5MW
0.5 N80-2.5MW
Capacity Factor

0.4

0.3

0.2

N90-2.3MW
0.1

0.0
3 6 9 12
Mean wind speed (m/s)
Table 3.4 Average wind turbine capacity
Tab le 1 0.4 Av erag e w ind tu rbin e
factors in 2001. Source: BTM Consult
cap acity factors in 2 001 . S ource:
(2002).
B T M C onsult (20 02).
C ou ntry C apacity F actor
UK 0.32
G re ece 0.29
D en m ark 0.26
S pa in 0.24
N eth erlan ds 0.24
C hin a 0.24
S w ed en 0.24
Italy 0.23
G erm any 0.21
India 0.20
Figure 3.20 Mean wind speed over North America at a height of 100 m.

3 4 5 6 7 8 9 10

Wind speed (m/s)


Source for this and other wind maps: Prepared from data file at power.larc.nasa.gov (go to Sustainable Buildings, Global Datasets)
Figure 3.21 Mean wind speed over Europe at a height of 100 m.

70

60

50

40

30
-10 10 30 50 70

3 4 5 6 7 8 9 10

Wind speed (m/s)


Figure 3.22 Mean wind speed over China and surrounding
regions at a height of 100 m.
54

44

34

24

14
70 90 110 130 150

2 3 4 5 6 7 8 9 1 0

Wind Speed (m/s)


Supplemental Figure: Mean wind speed over North Africa
and the Middle East at a height of 100 m.
45

30

15

0
-20 -5 10 25 40 55 70

0 2 3 4 5 6 7 8 9 10

Wind speed (m/s)


Supplemental figure: Mean wind speed over southern Africa
at a height of 100 m.
0

-6

-12

-18

-24

-30

-36
5 15 25 35 45

2 3 4 5 6 7 8

Wind speed (m/s)


Supplemental figure: Mean wind speed over Australia, Indonesia
and adjoining regions at a height of 100 m.
15

-1 5

-3 0
-32

-40

-4 5
95 115 135 155
-48
165 180

2 3 4 5 6 7 8

Wind speed (m/s)


15

-5

Supplemental
-15 figure: Mean
wind speed
-25 over South
America
-35
at a height of
100 m.
-45

-55
-83 -73 -63 -53 -43 -33

2 3 4 5 6 7 8 9 10

Wind speed (m/s)


Windfarms

• Clustering of many wind turbines in a regular


array in a region of good winds
• Turbines are typically spaced 5-9 rotor
diameters apart in the along-wind direction, and
3-5 rotor diameters apart in the cross-wind
direction
• Clustering reduces costs (economies of scale for
installation), and takes better advantage of the
best wind sites
Impact of windfarms on weather
and climate
• Large wind farms (involving hundreds of wind
turbines at the closest permitted spacing (i.e.,
separated by ~ 7 rotor diameters) would have a
noticeable effect on regional winds and hence
on vertical fluxes of heat and moisture in the
atmosphere, thereby changing the surface air
temperature in the region of the wind farm and
downstream from the wind farm
• Interference with the winds might reduce the
overall power output from a wind farm by up to
30% compared to the case where the winds are
assumed to be unaffected by the wind farm
Horizontal (left) and vertical (right) axis turbines

Source: Islam et al. (2013, RSER 21:456-468)


Vortices shed from the front row of fish assist the swimming of
fish in the next row, and so on. The same concept could be
applied to vertical-axis wind turbines – so the output from a
windfarm would be enhanced rather than diminished

Source: Islam et al. (2013, RSER 21:456-468)


Offshore wind farms
• Wind turbines mounted on the seabed in water
up to 50 m deep
• Can double or triple the cost of the wind turbine
+ connections to the grid, but there can easily be
twice the electricity production
• Net result – electricity for about the same to 50%
higher price but with twice the capacity factor
(i.e., 40-50% instead of 20-25%)
• Turbines especially designed for offshore
conditions have been built
Figure 3.25 Middelgrunden wind farm, next to Copenhagen

Source: Danny Harvey


Figure 3.26 European offshore wind atlas

Source: Risø National Laboratory, Wind Atlases of the World (www.windatlas.dk)


Figure 3.27 Types of foundations

G ra v ity b a sed su p p o r t str u c tu re

M o n o p ile su p p o rt stru ctu re


Tr ip o d su p p o rt str u ctu re

Source: Soker et al (2000, Offshore Wind Energy in the North Sea: Technical Possibilities and Ecological Considerations
- A Study for Greenpeace)
Growth in cumulative global offshore wind capacity, and
annual rate of addition, 2000-2016
Distribution of offshore wind capacity at the end of 2016

Source: Global Wind Energy Report 2016


Location and layout of the 576-MW Gwynt y Mor offshore
wind farm, under construction in 2015

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


Type of voltage of connections between European
offshore wind farms and the shore

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


Belgium offshore wind farms as of 2014 (2244 MW total)

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


Two North Sea wind farm network concepts

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


One of several US east coast (left) and east Asian (right)
concepts

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


Floating wind turbines
• There are about 30 different floating wind turbine concepts
under development, about 10 of which might reach full-scale
demonstration and fewer still reaching commercial deployment
• Floating wind turbines would need to be moored (attached) to
the sea bed with cables, so this does place some restrictions
on the water depth and seabed conditions where they can be
deployed
• 80-120 m will probably be the economically optimal depth for
floating wind turbines
• Potential wind energy capacity from floating wind farms is
nevertheless enormous: 4000 GW in Europe, 2450 in the US,
and 500 GW in Japan (recall: total global generation capacity
today of all types is around 6000 GW)
Countries where different floating wind turbine concepts are currently
under development – 6 different concepts are being worked on in
Japan, 5 in the US, 4 each in France and Norway

Source: Carbon Trust (2015, Floating Offshore Wind: Market and Technology Review)
Floating wind turbines

• One concept (WindSea) involves a triangular-


shaped floating platform with a 3.2MW turbine at
each corner on an outward-inclined tower
• One rotor would be on an airfoil-shaper tower
that would act like the tail of an airplane, serving to
continuously and automatically orient all three
rotors perpendicular to the wind
A floating wind turbine prototype.

Source: The Guradian, http://www.theguardian.com/environment/2014/jun/23/drifting-off-the-coast-of-portugal-the-


frontrunner-in-the-global-race-for-floating-windfarms?CMP=twt_gu
Fukushima floating wind turbine platform

Source: GWEC – Annual Market Update 2014


7-MW floating wind turbines under construction near
Fukushima, Japan, June 2015

urce: http://environews.tv/world-news/worlds-largest-offshore-wind-turbine-comes-online-near-fukushima-da
First commercial floating turbines, installed off the NE coast
of Scotland in 2017

Source: The Guardian, https://www.ft.com/content/2050b1de-72bc-11e7-aca6-c6bd07df1a3c)


Capacity Factors

Offshore areas tend to have stronger winds than onshore


winds, so going to offshore wind turbines but close enough to
the shore to be mounted on the seabed, there is a big
improvement. By being able to place wind turbines further
from shore, winds tend to be even stronger. Thus, capacity
factors will typically be larger:
• Onshore, 25-30% typical capacity factor
• Fixed-bed offshore, 40% typical
• Floating offshore, 50% typical

Double the capacity factor means that the capital and


maintenance costs can be doubled and the unit price of
electricity will be the same (but the supply will be more
reliable)
A network of floating offshore wind farms proposed for the North Sea

From C. Macilwain (2010, ‘Supergrid’, Nature 468, 624-625)


Costs: highly uncertain, but expected to fall a lot – just as the
investments required to expand oil production in the North Sea fell
greatly with the transition from fixed to floating oil rigs, as shown below.

Source: Carbon Trust (2015, Floating Offshore Wind: Market and Technology Review)
Expertise from the offshore oil and gas industry that is
useful for the offshore wind energy:
• Experience with floating platforms
• Experience with autonomous underwater robots for repairs and
maintenance
• Experience in managing financial, political and project-development risks
• Expertise with network connectivity between oil and gas platforms and
nearby ships

The port infrastructure developed to serve the offshore oil and gas industry
(in Europe) could be used to service an offshore wind energy industry.

The developing offshore wind energy industry offers something for oil and
gas companies to transition to for the post-fossil fuel era – rather than
going out of business.
Fluctuations in Wind Electricity Production

• Because there might be times when the wind speed might be less
than the cut-in wind speed (so that no electricity is produced), some
amount of non-wind backup capacity is needed (how much will be
discussed later)
• As well, because wind is variable, some power units that can go up
and down to offset the variations in wind electricity production are
needed
• The problem is, the units most able to fluctuate rapidly (such as
simple-cycle natural gas turbines) tend to be less efficient than the
units that would normally be used (such as coal steam turbines or
combined-cycle natural gas systems)
• Thus, it is desirable to minimize the variations in the electricity
production from wind
Supplemental Slide: Impact on GHG emissions when wind replaces
electricity from coal in Illinois – there are increased emissions from the
natural gas powerplants needed to balance variations in wind energy,
especially as less efficient single-cycle (but fast responding) turbines
are needed part of the time, and this offsets about 15% of the emission
reduction that would otherwise occur when wind supplies 40% of the
electricity otherwise supplied by coal.

(from natural gas plants)

Source: Valentino et al (2013, Env Sci. & Technology 46:4200-4206)


Minimizing rapid (seconds to minutes)
fluctuations in wind output

• Use of variable-speed turbines – provides some


smoothing of output on a time scale of seconds
• Link together several turbines in a wind farm –
provides some cancellation of fluctuations at a time
scale of up to a minute or so
• Implement short-term storage of excess energy
- flywheels, supercapacitors, plug-in hybrid vehicles
with a two-way connection to the grid
Dealing with longer fluctuations
(hours to days and months)

• Link together wind farms over a broad region


• Use electrolyzers and fuels cells (making and using
H 2)
• Use flow batteries (regenerative fuel cells)
• Use underground compressed air energy storage
(CAES)
• Use hydro-electric reservoirs
• Use heat pumps and thermal energy storage in
district energy systems
• Use other flexible end-use electric loads
Table 3.10 Impact on the statistical properties of wind
energy of spreading wind farms over increasingly
larger areas in and around Europe

Source: Czisch and Giebel (2000, Wind Power for the 21st Century, Kassel)
Figure 3.29 Amalgamation of dispersed wind farms

1.0 P M ean /P R ated 1.5


0.37 No rthern Russ ia and W estern Siberia
M o nthly M ean W ind Tu rb ine - Po w er

0.9

M o nthly M ean Electric D em and


0.28 Kazakhstan
0.8 0.38 S ou thern Mo rocco 1.2
0.36 M auritania
0.7 0.30 Go od W ind S ites within EU and Norway
0.33 Co m bination : 1/3 e) and each 1/8 a), b), c) and d)

(P M ean /P R ate d )
0.6 0.47 E lectric De m and w ithin EU and N orw ay 0.9
(P M ean /P R ated )

0.5

0.4 0.6

0.3

0.2 0.3

0.1

0.0 0.0
1 2 3 4 5 6 7 8 9 10 11 12
Tim e (M o nth)
Source: Czisch and Giebel (2000, Wind Power for the 21st Century, Kassel)
Table 3.11 Largest variation in wind power over different
time periods and averaged over differently-sized
regions.
Largest
Areal dimensions variation up Location
or down
Hourly variations
100 x 100 km 50% UK
200 x 200 km 30% Denmark
400 x 400 km 20% Germany, Denmark,
Finland
Group of countries 10%
4-12 hour variations
One country 40-60% Denmark
80% Germany
Larger area 35% Nordic area
400 x 400 km 4 hr: 80%
6 hr: 80% UK
12 hr: 90%
Source: EWEA (2005, Large-scale Integration of Wind Energy in the European Power Supply,
www.ewea.org)
Making Use of Short-term Wind
Forecasts
• The variability in electricity output that remains
after making use of the various strategies
outlined in the previous slides can be better
handled if the variation can be predicted several
hours in advance, as this permits scheduling of
slowly responding backup fossil fuel power units
• Thus, improving local wind forecasts with high-
resolution meteorological computer forecasting
models is an intensive area of research at
present
Electrolyzers and fuel cells
• Electrolyzers generate hydrogen by splitting water
using electricity, and so could use excess wind
electricity
• The hydrogen can be stored as a compressed gas
• When there is a shortage of wind electricity, additional
electricity can be generated by running the electrolyzer
backwards as a fuel cell
• Rapid variations in output degrade the performance
and shorten the lifespan of electrolyzers and fuel cells,
so a battery would likely be used to smooth out the
electricity input to the electrolyzer and smooth the
demand for extra electricity from the fuel cell
Compressed air energy storage (CAES)
• With electricity generation using a gas turbine, about
half of the turbine power is used to compress the air
needed for combustion, and only about half is used to
drive the generator that produces electricity
• If excess wind energy is used to compress air and store
it underground, the compressed air could be directly
used in a gas turbine to generate electricity when there
is a shortage of wind power
• This would more than double the efficiency of using
natural gas to produce electricity – from about 37%
(with a simple-cycle turbine) to 84%
Geological formations
suitable for CAES

• Salt domes, salt beds and porous sedimentary


rocks are best
• These underlie 75% of the land area of the US,
including many of the best wind regions
• Salt domes closely coincide with the best wind
resource regions in Europe
• Caverns can also be excavated in hard rock, but
these would be considerably more expensive
Supplemental Figure: Location of geology suitable for CAES
and of good wind resources in the US, and CAES sites.

Source: Succar and Williams (2008, Compressed Air Energy Storage: Theory, Resources, and Applications for
Wind Power, Princeton Environment Institute, Princeton University, Princeton, New Jersey)
Figure 3.31 Wind CAES energy flow

Fuel C O 2
100 M J 1.4 kgC
(80-120) (1.1-1.7)
“Spilled” C AES From CA ES
8 kW h ER =0.74 22 kW h
(5-15) HR =4649 (18-27)
kJ/kW h
To C A ES
16 kW h
(13-20) To Load
78 kW h
(73-82)

W ind Input Firm Total O utput


102 kW h 100 kW h
(97-110)
Source: Denholm (2006, Renewable Energy 31, 1355–1370, http://www.sciencedirect.com/science/journal/09601481)
CAES was initially developed in the 1970s and
1980s as a means of absorbing excess nuclear
electricity (the output from a nuclear powerplant is
largely fixed, while electricity demand varies, so if
the plant meets a large fraction of peak demand,
there would be excess power at times)

Only two such plants were built (one in Germany,


one in Alabama)

Now there is a revival of CAES, with many plants


under construction or planned (especially in
Texas) for storage of excess wind energy
Existing and planned CAES facilities require some
supplemental fuel (natural gas at present, but it
could be gasified biomass in the future)

A new system under development is called


advanced adiabatic CAES (AA-CAES)

In this system, no (or very little) supplemental fuel


is required. Instead, the heat that is produced
when air is compressed is stored in the form of hot
(~ 650ºC) molten salts in an insulated tank, and
used instead of fuel along with the compressed air
to generate electricity when needed
Use of hydro-electric reservoirs
• Most hydro-electric reservoirs are not running at full
capacity all the time, because there is not enough water
• Thus, when there is excess wind, the water flow and
hence electricity production can be reduced, and when
there is a shortage, greater water flow than would
otherwise be the case (using the saved water) can be
allowed
• This entails no energy loss, and in fact can slightly
increase the annual hydro-electric energy production
from the same annual water flow, because the average
reservoir level will be greater (hydro-electric power
production depends on flow rate x elevation drop)
• Many areas of the world that are lousy for CAES
(such as the hard pre-Cambrian rocks of the
Canadian Shield) have excellent hydro-electric
resources or excellent potential hydro-electric
resources
Pumped hydro

• When there is no reservoir to hold back the river


flow, build a dam in some mountain valley and
pump water up behind the dam using excess
wind electricity, creating a reservoir
• Let water drain the reservoir through a
conventional hydro-electric turbine at times of
wind electricity deficit
• This is used in Europe (in the Alps, for example)
Source: Renewable Energy World, March-April 2017, p. 12
Very small pumped hydro storage within the wind turbine tower

Source: Renewable Energy World, March-April 2017, p. 14


Use of heat pumps
• One way to provide heat with electricity is through electric
resistance heating, but this is not efficient (only 1 unit of heat
per unit of electricity used)
• A better method is to use electric heat pumps, which provide
3-4 units of heat per unit of electricity used
• If we have well-insulated buildings (such that they can drift for
a few hours without the heating system on), then heat pumps
could be used when there is excess wind and the heating
turned off altogether when wind energy drops
• Similarly, heat pumps (or chillers) can be used for cooling
purposes (in the summer) at times of excess wind and turned
off at times of deficit
• In effect, excess wind energy is being stored as thermal
energy (heat or coldness)
• There is no loss of energy in this way, so the
storage efficiency is 100%
• If, however, heat or coldness is stored in a large
insulated tank outside the building (as in some
district energy systems), there would be some
loss of stored heat or coldness to the
environment
• The storage efficiency in this case is < 100%,
perhaps 95%
Other dispatchable loads
using dynamic demand
• The power industry talks about “dispatchable”
power sources – those that can be quickly
brought on line and varied in output to meet
fluctuating electricity demand
• The other side is to have “dispatchable” demand
– demand that can be reduced by the power
utility to compensate for surges in demand
elsewhere or for the loss of power units
• This is already done with things like electric
resistance water heaters – a signal can be sent
from the utility to temporarily turn them off
• When electricity demand exceeds electricity
supply, the voltage and frequency drop (and vice
versa when supply exceeds demand)
• Equipment with compressors (such as
refrigerators and air conditioners) can sense
changes in frequency and can now be designed
to automatically shut down when the frequency
drops below some threshold
• This could compensate for sudden drops in wind
power until the wind power resumes or backup
systems come on line
• This is called dynamic demand
Present round-trip efficiency of various energy
storage option (energy taken out vs energy put in)
• Electrolyzer/hydrogen/fuel cell system: 32-42%
• Flywheels (store for 1 day): 45%
• Pumped hydro: 65-80%
• Flow batteries: 75%
• CAES: 70-75%
• AA-CAES: 70%
• Flywheels (store for seconds): 85%
• Batteries: 85-90%
• Capacitors: 95%
• Thermal energy storage: 95-100%
• Hydro-electric reservoirs: ≥ 100%
Improving system stability through the
addition of wind turbines
• Until recently, wind turbines had been thought of only as a liability as far as
system stability is concerned (due to the fluctuating and partly
unpredictable nature of wind)
• However, modern wind turbines can be used to improve overall system
stability by
- compensating for shifts in reactive power (related to the phase shift
between voltage and current oscillations) caused by other supply sources
or loads in the system (some turbines can provide reactive power even
when not spinning!) (REW Jan-Feb 2017)
- maintaining connection to the grid and continuing to produce power when
faults elsewhere cause large transient variations in voltage (this is called
fault ride through (FRT) capability, and is still under development)
- varying their output in response to changes in grid frequency (this comes
with a cost – output has to be restrained slightly under normal conditions)
Transmission
Transmission basics

• Transmitted power = Voltage (V) x Current (I)


• Resistance loss varies with I2, so,
• For a given energy flow, the resistance loss varies with
1/V2
• Thus, the key to minimizing resistance losses is to
transmit electricity at high voltage
• There were be some offsetting losses in the transformers
from low to high and back to low voltage
• Typical voltages for long distance transmission:
500-800 kV, compared to 30 kV for local distribution
HVDC (high voltage DC)
• Less expensive transmission lines with smaller resistance losses, but more expensive
transformers with greater losses
• If the wind turbine produces AC electricity (as in the popular DFIG type), then we’d
need to convert from AC to DC with a transformer (entailing some energy loss),
transmit as DC, and then convert back to AC

• However, HVDC costs less and entails less overall loss for transmission beyond some
minimum distance, namely,
• HVDC costs less beyond about 750 km distance, and entails less loss beyond about
250 km distance (the exact break-even distance for cost depends on the terrain and
local market conditions)

• If the wind turbine directly produces DC electricity as a first step (as in the PMSG),
then we could save on DC-to-AC conversion that is otherwise done in these turbines,
and avoid the need for and AC/DC transformer prior to the HVDC transmission
Recall: Permanent magnet synchronous generator (may or
may not have gears, allows full variation in turbine rotor speed)
Other pros and cons of HVDC
compared to HVAC
• HVDC
- has a much narrower right of way
- generates negligible magnetic fields (concern over which has been one
source of public opposition to new transmission lines)
- has better reactive power control and full control of where the power
flows (unlike AC mesh grids)
- an offshore grid for offshore wind farms would permit the wind turbines
to operate at a greater range of speeds, which in turn would permit
more efficient operation (no need for synchronization of the power
output with the land AC grid)
• However,
- branching of DC lines is difficult, as is the construction of multiple
terminals, although these problem should be solvable in a few years
Figure 3.32 Typical DC and AC Transmission Pylons

± 500 kV DC 800 kV AC
route width: 50 m 85 m

Source: GAC (2006, Trans-Mediterranean Interconnection for Concentrating Solar Power, Final Report, GAC,
www.dlr.de/tt/trans-csp)
Figure 3.33 Transmission corridors transmitting
10 GW of electric power

Source: GAC (2006, Trans-Mediterranean Interconnection for Concentrating Solar Power, Final Report, GAC,
www.dlr.de/tt/trans-csp)
Cost of electricity from wind
Direct Cost of Wind Energy

The cost per kWh is the annual revenue requirement per


kW of capacity divided by the number of kWhs sold per
year per kW of capacity. That is,

C = (CRF+OM)*CCwt/(ηsfa 8760*CF)

where CRF=cost recovery factor


= i/(1-(1+i)(-N))
i = interest rate (expressed as a fraction per year)
N = number of years over which the wind project is
financed
OM = annual operation and maintenance cost as a
fraction of the initial capital cost CCwt,
CCwt = initial capital cost given as $/kW ($ per kW of
turbine capacity)
8760 = number of hours in a year
ηs is an efficiency that takes into account various losses
that are not accounted for in the turbine power curve (such
as dirt on the blades, imperfect tracking of the wind
direction by the yaw mechanism, or wake effects in wind
farms)
fa is the fraction of time that the turbine is available
CF = capacity factor (the average power output as a
fraction of the peak output or capacity)

a 1kW turbine running full out all the time would produce
1kW x 8760 hr/yr = 8760 kWh/yr of electricity
Units in the previous equation:

(yr)-1 (for OM and CRF) x $/kW, divided by

kWh/kW/yr

gives

$/kWh
Figure 3.36: Illustrative costs of wind electricity for various rates
of return (ROI) in the investment and for various capital costs,
assuming a CF of 0.35, 20-year financing andηs = fa = 1.0
16

14 12%/yr ROI
Electricity Cost (cents/kWh)

12

10 6%/yr ROI

8 3%/yr ROI

0
600 800 1000 1200 1400 1600 1800 2000 2200 2400
Capital Cost ($/kW)
Cost of solar and wind energy contracted to begin in 2016-2019 (divide
by 10 to get cents/kWh)

Source: IEA (2016, Energy, Climate Change, and Environment: 2016 Insights)
Some wind costs (US cents)
from the preceding slide:

Germany, 6.7-10.0 cents/kWh


China, 8.0-9.1 cents/kWh
Canada, 6.6 cents/kWh
Egypt, 4.1-5.0 cents/kWh
Peru, 3.8 cents/kWh
Morocco, 3.0-3.5 cents/kWh
Indirect costs of wind turbines

• Reduced electricity output by non-wind


generators (which increases the unit cost of their
electricity), partly offset by a reduction in the
need for non-wind generators
• Wasted wind electricity generation potential
It is often thought that the addition of wind energy
does not allow any reduction in the amount
(capacity) of other power sources, because there
could be zero wind production near times of peak
demand. That is, the capacity credit of wind is
often assumed to be zero. However, this is not
correct.

Instead, the amount of non-wind capacity that is


needed is calculated so as to have the same loss-
of-load probability as when there is no wind
capacity with, instead, the full non-wind capacity
The result is that, for small wind penetration (that
is, small wind capacity compared to the total
capacity), the capacity credit is roughly equal to
the capacity factor.

So, if 100 MW of wind power capacity is added to


a very large system and the capacity factor
(average output as a fraction of peak output) of the
wind turbines is 20%, then the non-wind capacity
can be reduced by 20 MW while still having the
same overall system reliability.
As the wind penetration increases,
the capacity credit as a fraction of the capacity
factor becomes progressively smaller

So, at 1000 MW and the same 20% capacity


factor, the capacity credit might be only 10%
instead of 20%, so the non-wind capacity can be
reduced by only 100 MW while having the same
overall system reliability

The capacity credit for wind is non-zero only


because the backup fossil fuel powerplants are
themselves not 100% reliable, as seen in the
next table.
Table 3.16 Outage rates for various electricity
generators in the US.

G en era to r Ty p e O u tra g e R a te (% )
F o rced P la n n ed
H y d ro 2 .0 5 .0
G a s tu rb in e 1 0 .7 6 .4
G a s co m b in ed cy c le 5 .0 7 .0
E x istin g c o a l 7 .9 9 .8
N ew coal 7 .9 9 .8
IG C C 7 .9 9 .8
N u clea r 5 .0 5 .0
Figure 3.38 Capacity credit for wind as a function of the
wind penetration and capacity factor (CF)
50
Capacity Credit (% of Wind Capacity)
CFwind=0.40
40
CFwind=0.30

30 CFwind=0.20

CFwind=0.10
20

10

0
0 10 20 30 40 50
Wind Penetration (% of System Peak)
Figure 3.39 Capacity credit for wind as a function of
wind penetration and the degree of geographical
dispersion of the wind turbines
Capacity Credit (% of Wind Capacity) 50

Perfect dispersion, d=0.0


40

d=0.2

30
d=0.5

20

10

Perfection concentration, d=1.0

0
0 10 20 30 40 50
Wind Penetration (% of System Peak)
To recap,
• The addition of wind means that the existing fossil fuel
powerplant is used less, which increases the unit cost
of that portion of the electricity from the fossil fuel plant
• However, less fossil fuel powerplant is needed (which
is not helpful if the fossil capacity has already been
built) due to the non-zero capacity credit from wind
• Other indirect costs of wind include
- wasted wind electricity potential due to the need to
maintain a minimum fossil fuel output
- reduction in the efficiency of the fossil fuel powerplant
when wind is added (either because it is operating at
lower average load and thus less efficiently, or
because of larger swings in output)
Figure 3.40 Wasted wind energy potential as a function of
wind energy penetration for Danish conditions

100
% of electricity demand

80 Demand not met by Wind

60

40
Wasted Wind Energy
20

0
0 20 40 60 80 100

Wind energy production as a % of electricity demand

Source: Redlinger et al (2002, Wind Energy in the 21st Century: Economics, Policy, Technology and the
Changing Electricity Industry, Palgrave, Basingstoke)
The preceding figure shows that, in Denmark, if
enough wind turbines were installed to be able to
generate an amount of electricity in one year equal
to total annual electricity demand (i.e., 100% wind
energy penetration), in fact only 60% of the
electricity demand would be met by wind, and 40%
of the generated wind electricity would be wasted
due to mismatches between fluctuating supply and
demand. So – ways to adjust demand or store
excess electricity would be needed.
Trends in the capital cost of
onshore wind turbines ($/kW)
and in the cost of electricity
from wind ($/kWh)
Capital cost of wind turbines:
US and global average trends

Source: Wiser and Bolinger (2016, 2015 Wind Technologies Market Report)
Variation of wind farm unit cost with the size of the
wind farm

Source: Wiser and Bolinger (2016, 2015 Wind Technologies Market Report)
Variation of wind farm unit cost with the size of the
wind turbines used

Source: Wiser and Bolinger (2016, 2015 Wind Technologies Market Report)
The purchase price of turbines has been reduced
by up to 45% from the list price for orders of 500-
1600 turbines
Declining cost of wind energy in the US (divide by 10 to get cents/kWh)
(a decrease from 10-17 cents/kWh in 2009 to 3.2-6.2 cents/kWh by 2016)

Source: ACORE (2017, The State of Play of Renewable Energy)


Projections of future decreases in the cost of wind energy in the US with
accelerated research and development, for a given wind regime
(absolute costs will vary regionally with labour and transport costs)

Costs are projected to drop from 4.3-5.0 cents/kWh to 2.0-3.5 cents/kWh

Source: Dykes et al. (2017, NREL Science-based innovation report)


Factors contributing to the further 30-60%
cost-of-electricity reduction

• Reduction in $/kW capital cost by ~ 20% through scaling up


the size of the turbine, and various innovations in design and
manufacturing
• Increase in electricity production by ~ 20% through
innovations in turbine design and development of “smart” wind
farms
• Reducing financing costs by 5% by reducing the perceived
risk to investors
• Reducing cost recovery factor by ~ 25% by extending wind
turbine lifespan and hence amortization period to 30 years
“Smart” wind farms
• Developing the ability to model the 3-D airflow through
an entire wind farm, not just around a single, isolated
wind turbine
• Using the above to optimize of the size of each
individual wind turbine within a wind farm
• Development of techniques to sense the real air flow
through the wind farm
• Development of techniques to actively modulate the
airflow through a wind farm (be adjusting the blades of
individual turbines) so as to maximize the total wind
farm electricity production
Cost of offshore wind turbines
(fixed to the ocean bed
and floating)
Cost of European offshore wind projects – some high capital
cost projects (up to 7000 euro/kW), but capacity factors are
40-50%

Source: Rodrigues et al (2015, Renew. Sust. Energy Rev. 49:1114-1135)


Cost of electricity from European offshore wind projects

Costs are 7.5-20 cents/kWh, with lower costs projected for the
latest projects (to be completed by 2021-23)

Source: Dykes et al. (2017, NREL Science-based innovation report)


Current and projected cost of electricity from offshore wind:
16 cents (US)/kWh, dropping to 11 cents/kWh
(onshore wind is generally 4-8 cents/kWh)

Source: Renewable Energy World, Nov-Dec 2016


The next slide gives the estimated capital
cost (per MW) for different floating wind
turbine concepts for a 500 MW wind farm
consisting of 100 5-MW wind turbines in
water 200 m deep, 200 km from shore. Note
that mooring and grid costs account for
about half of the total cost.
4.5 million euro/MW corresponds to 4500
euro/kW, so the range of capital costs is
roughly US$4000-5000/kW
Source: Myhr et al. (2014 , Renewable Energy 66:714-728)
Variation in cost of electricity with water depth and distance to
shore for different floating wind turbine concepts. The extreme
cost range is 14-23 eurocents/kWh or 17-27 cents/kWh

Source: Myhr et al. (2014, Renewable Energy 66:714-728)


Trends in the performance of
wind turbines
Improvement in the capacity factor of successive generations
of wind turbines (from 2002-03 to 2012-3) for a given wind
speed

Source: Islam et al. (2013, RSER 21:456-468)


Capacity factor of wind turbines in the US, by year of installation

Source: IEA (2016, Next Generation Wind and Solar Power)


Trend in average turbine capacity, rotor diameter, and hub height in the
US based on the year the wind turbine was installed
From 2011, capacity hardly changed but rotors got larger, making it easier to start up
in low-wind conditions and thereby increasing the capacity factor

Source: Wiser and Bolinger (2016, 2015 Wind Technologies Market Report)
Variation of wind capacity factor in 2015 according to the year the wind
turbine was installed (blue bars) and of possible explanatory factors
(resource quality, hub height, and rotor size-to-generator capacity ratio)
Note that capacity factor improved from 25% for pre-2000 turbines to 41% for
turbines installed in 2014, although winds were slightly worse at the 2014 sites.

Source: Wiser and Bolinger (2016, 2015 Wind Technologies Market Report)
Comparison of power curve of recent and latest wind turbines.

Depending on the frequency of different wind speeds, the advanced turbine


could have twice the capacity factor of the classical turbine, at a given
location. A big factor in this is the use of larger rotors relative to the
generator capacity.

Source: Hirth and Muller (2016)


Generation profiles of older and newer wind turbines in a
given location – reduced maximum electrical power but larger
minima – resulting in a smoother and less extreme variation.

Source: IEA (2016, Next Generation Wind and Solar Power)


Strategies for Baseload Wind Energy

• Oversized wind farms compared to the


transmission link – can give capacity factors at
the receiving end of the link of 0.6-0.7
• Compressed air energy storage
• Use of dispatchable loads (such as reverse
osmosis for desalination or heat pumps in
district heating and cooling systems)
Figure 3.44 Oversizing Concept
Figure 3.45c Transmission line capacity factor as a function of
mean wind speed for various degrees of wind farm over-sizing

0.8
4
3
2
Transmission Capacity Factor

0.6
1

Oversizing factor
0.4

0.2

0
4 6 8 10 12

Mean Wind Speed (m/s)


Capital Cost Estimates

• Natural gas combined cycle: $660/kW


• Wind – natural gas hybrid: $1640/kW
• Wind – natural gas – CAES: $2270/kW
(the last is cheaper than many recent
nuclear power plants)
Energy Payback Time

• This is the time required for the amount of


primary energy saved by the wind turbine to
offset the total primary required to produce the
turbine
• Saved primary energy per year = electrical
energy produced per year divided by the
efficiency of the powerplant that would otherwise
be used to produce electricity
• Generally speaking, calculated payback
times for wind turbines are 2-8 months
• The payback time would be significantly
longer if the components need to be
transported 1000 km or more by truck
Noise and impacts on birds and bats

• Bird mortality is miniscule compared to many


other human causes of bird deaths (see table to
follow)
• Noise level at a distance of 350 m is less than
the typical background level in a home
• Impacts on bats need further study
Table 3.24 Main human-related causes of bird
deaths in the US.

C a u se E stim a ted n u m b er o f
d eath s p er y ear
U tility tra n sm ission an d d istrib u tio n lin es 13 0-17 4 m illion
C o llision s w ith road veh icles 60-80 m illion
C o llision s w ith b u ild in g s 10 0-1,0 00 m illio n
Telecom m u n ica tio n s tow ers 40-50 m illion
A g ricu ltu ral p esticid es 6 7 m illion
C a ts 3 9 m illion

Source: GWEC (2006, Global Wind Energy Outlook 2006, www.gwec.net)


How much wind energy capacity
would we need globally as part
of a C-free, 100% renewable
energy system?

You might also like