You are on page 1of 45

Wafer (electronics)

In electronics, a wafer (also called a slice


or substrate)[1] is a thin slice of
semiconductor, such as a crystalline
silicon (c-Si), used for the fabrication of
integrated circuits and, in photovoltaics,
to manufacture solar cells. The wafer
serves as the substrate for
microelectronic devices built in and upon
the wafer. It undergoes many
microfabrication processes, such as
doping, ion implantation, etching, thin-
film deposition of various materials, and
photolithographic patterning. Finally, the
individual microcircuits are separated by
wafer dicing and packaged as an
integrated circuit.
Top: polished 12" and 6" silicon wafers.
Their crystallographic orientation is
marked by notches and flat cuts (left).
VLSI microcircuits fabricated on a 12-inch
(300 mm) silicon wafer, before dicing and
packaging (right).
Bottom: solar wafers on the conveyor
(left) and completed solar wafer (right)

History
In the semiconductor industry, the term
wafer appeared in the 1950s to describe
a thin round slice of semiconductor
material, typically germanium or silicon.
Round shape comes from single-crystal
ingots usually produced using
Czochralski process. Silicon wafers were
first introduced in the 1940s.[2][3]

In the 1950s, Mohamed Atalla


investigated the surface properties of
silicon semiconductors at Bell Labs,
where he adopted a new method of
semiconductor device fabrication,
coating a silicon wafer with an insulating
layer of silicon oxide, so that electricity
could reliably penetrate to the conducting
silicon below, overcoming the surface
states that prevented electricity from
reaching the semiconducting layer. This
is known as surface passivation, a
method that later became critical to the
semiconductor industry as it made
possible the mass-production of silicon
integrated circuits (ICs).[4][5][6] The
surface passivation method was
presented by Atalla in 1957,[7] and was
later the basis for the metal-oxide-
semiconductor (MOS) process invented
by Atalla and Dawon Kahng in 1959.[4]

By 1960, silicon wafers were being


manufactured in the U.S. by companies
such as MEMC/SunEdison. In 1965,
American engineers Eric O. Ernst, Donald
J. Hurd, and Gerard Seeley, while working
under IBM, filed Patent US3423629A[8]
for the first high-capacity epitaxial
apparatus.

Silicon wafers are made by companies


such as Sumco, Shin-Etsu Chemical,[9]
Hemlock Semiconductor Corporation
and Siltronic.

Formation

The Czochralski process.


Wafers are formed of highly pure,[10]
nearly defect-free single crystalline
material, with a purity of 99.9999999%
(9N) or higher.[11] One process for
forming crystalline wafers is known as
Czochralski growth invented by the
Polish chemist Jan Czochralski. In this
process, a cylindrical ingot of high purity
monocrystalline semiconductor, such as
silicon or germanium, called a boule, is
formed by pulling a seed crystal from a
melt.[12][13] Donor impurity atoms, such
as boron or phosphorus in the case of
silicon, can be added to the molten
intrinsic material in precise amounts in
order to dope the crystal, thus changing
it into an extrinsic semiconductor of n-
type or p-type.

The boule is then sliced with a wafer saw


(a type of wire saw) and polished to form
wafers.[14] The size of wafers for
photovoltaics is 100–200 mm square
and the thickness is 100–500 μm.[15]
Electronics use wafer sizes from 100–
450 mm diameter. The largest wafers
made have a diameter of 450 mm[16] but
are not yet in general use.

Cleaning, texturing and


etching
Wafers are cleaned with weak acids to
remove unwanted particles, or repair
damage caused during the sawing
process. There are several standard
cleaning procedures to make sure the
surface of a silicon wafer contains no
contamination. One of the most effective
methods is RCA clean. When used for
solar cells, the wafers are textured to
create a rough surface to increase their
efficiency. The generated PSG
(phosphosilicate glass) is removed from
the edge of the wafer in the etching.[17]

Wafer properties

Standard wafer sizes …

Silicon …
Silicon wafers are available in a variety of
diameters from 25.4 mm (1 inch) to
300 mm (11.8 inches).[18][19]
Semiconductor fabrication plants,
colloquially known as fabs, are defined by
the diameter of wafers that they are
tooled to produce. The diameter has
gradually increased to improve
throughput and reduce cost with the
current state-of-the-art fab using
300 mm, with a proposal to adopt
450 mm.[20][21] Intel, TSMC and Samsung
were separately conducting research to
the advent of 450 mm "prototype"
(research) fabs, though serious hurdles
remain.
2-inch (51 mm), 4-inch (100 mm), 6-inch (150 mm),
and 8-inch (200 mm) wafers
Year
Typical Weight 100 mm2 (10 mm)
Wafer size introduced
Thickness [18]
per wafer Die per wafer

1-inch (25 mm) 1960

2-inch (51 mm) 275 μm 1969

3-inch (76 mm) 375 μm 1972

10 grams
4-inch (100 mm) 525 μm 1976 [22]
56

4.9 inch (125 mm) 625 μm 1981

150 mm (5.9 inch, usually


675 μm 1983
referred to as "6 inch")

200 mm (7.9 inch, usually 53 grams


725 μm. 1992 [22]
269
referred to as "8 inch")

300 mm (11.8 inch, usually 125


775 μm 2002 640
referred to as "12 inch") grams[22]

450 mm (17.7 inch) 342 grams


925 μm - 1490
(proposed)[23] [22]

675-millimetre (26.6 in)
Unknown. -
(theoretical)[24]

Wafers grown using materials other than


silicon will have different thicknesses
than a silicon wafer of the same
diameter. Wafer thickness is determined
by the mechanical strength of the
material used; the wafer must be thick
enough to support its own weight
without cracking during handling. The
tabulated thicknesses relate to when that
process was introduced, and are not
necessarily correct currently, for example
the IBM BiCMOS7WL process is on 8 in
wafers, but these are only 200um thick.
The weight of the wafer goes up along
with its thickness and diameter.

Historical increases of wafer size …

A unit wafer fabrication step, such as an


etch step, can produce more chips
proportional to the increase in wafer
area, while the cost of the unit fabrication
step goes up more slowly than the wafer
area. This was the cost basis for
increasing wafer size. Conversion to
300 mm wafers from 200 mm wafers
began in earnest in 2000, and reduced
the price per die about 30-40%.[25] Larger
diameter wafers allow for more die per
wafer.

Photovoltaic …

M1 wafer size (156.75 mm) is in process


of phase-out in China as of 2020. A
number of nonstandard sizes has arisen,
so efforts to produce M10 standard (182
mm) is an ongoing effort. Like
semiconductor cousin, driving down
costs is main driver despite that purity
requirements are completely different.
Proposed 450 mm transition …

There is considerable resistance to the


450 mm transition despite the possible
productivity improvement, because of
concern about insufficient return on
investment.[25] There are also issues
related to increased inter-die / edge-to-
edge wafer variation and additional edge
defects. 450mm wafers are expected to
cost 4 times as much as 300mm wafers,
and equipment costs are expected to rise
by 20 to 50%.[26] Higher cost
semiconductor fabrication equipment for
larger wafers increases the cost of
450 mm fabs (semiconductor fabrication
facilities or factories). Lithographer Chris
Mack claimed in 2012 that the overall
price per die for 450 mm wafers would
be reduced by only 10–20% compared to
300 mm wafers, because over 50% of
total wafer processing costs are
lithography-related. Converting to larger
450 mm wafers would reduce price per
die only for process operations such as
etch where cost is related to wafer count,
not wafer area. Cost for processes such
as lithography is proportional to wafer
area, and larger wafers would not reduce
the lithography contribution to die
cost.[27] Nikon planned to deliver 450-mm
lithography equipment in 2015, with
volume production in 2017.[28][29] In
November 2013 ASML paused
development of 450-mm lithography
equipment, citing uncertain timing of
chipmaker demand.[30]

The timeline for 450 mm has not been


fixed. In 2012, it was expected that
450mm production would start in 2017,
which never realized.[31][32] Mark Durcan,
then CEO of Micron Technology, said in
February 2014 that he expects 450 mm
adoption to be delayed indefinitely or
discontinued. "I am not convinced that
450mm will ever happen but, to the
extent that it does, it’s a long way out in
the future. There is not a lot of necessity
for Micron, at least over the next five
years, to be spending a lot of money on
450mm. There is a lot of investment that
needs to go on in the equipment
community to make that happen. And the
value at the end of the day – so that
customers would buy that equipment – I
think is dubious."[33] As of March 2014,
Intel Corporation expected 450 mm
deployment by 2020 (by the end of this
decade).[34] Mark LaPedus of
semiengineering.com reported in mid-
2014 that chipmakers had delayed
adoption of 450 mm "for the foreseeable
future." According to this report some
observers expected 2018 to 2020, while
G. Dan Hutcheson, chief executive of
VLSI Research, didn't see 450mm fabs
moving into production until 2020 to
2025.[35]

The step up to 300 mm required major


changes, with fully automated factories
using 300 mm wafers versus barely
automated factories for the 200 mm
wafers, partly because a FOUP for
300 mm wafers weighs about 7.5
kilograms[36] when loaded with 25
300 mm wafers where a SMIF weighs
about 4.8 kilograms[37][38][22] when
loaded with 25 200 mm wafers, thus
requiring twice the amount of physical
strength from factory workers, and
increasing fatigue. 300mm FOUPs have
handles so that they can be still be
moved by hand. 450mm FOUPs weigh 45
kilograms[39] when loaded with 25
450 mm wafers, thus cranes are
necessary to manually handle the
FOUPs[40] and handles are no longer
present in the FOUP. FOUPs are moved
around using material handling systems
from Muratec or Daifuku. These major
investments were undertaken in the
economic downturn following the dot-
com bubble, resulting in huge resistance
to upgrading to 450 mm by the original
timeframe. On the ramp-up to 450 mm,
the crystal ingots will be 3 times heavier
(total weight a metric ton) and take 2–4
times longer to cool, and the process
time will be double.[41] All told, the
development of 450 mm wafers requires
significant engineering, time, and cost to
overcome.

Analytical die count estimation …

In order to minimize the cost per die,


manufacturers wish to maximize the
number of dies that can be made from a
single wafer; dies always have a square
or rectangular shape due to the
constraint of wafer dicing. In general, this
is a computationally complex problem
with no analytical solution, dependent on
both the area of the dies as well as their
aspect ratio (square or rectangular) and
other considerations such as the width of
the scribeline or saw lane, and additional
space occupied by alignment and test
structures. Note that gross DPW
formulas account only for wafer area that
is lost because it cannot be used to
make physically complete dies; gross
DPW calculations do not account for
yield loss due to defects or parametric
issues.

Wafermap showing fully patterned dies, and partially


patterned dies which don't fully lie within the wafer.
Nevertheless, the number of gross die
per wafer (DPW) can be estimated
starting with the first-order
approximation or wafer-to-die area ratio,

where is the wafer diameter (typically


in mm) and the size of each die (mm2)
including the width of the scribeline ( or
in the case of a saw lane, the kerf plus a
tolerance). This formula simply states
that the number of dies which can fit on
the wafer cannot exceed the area of the
wafer divided by the area of each
individual die. It will always overestimate
the true best-case gross DPW, since it
includes the area of partially patterned
dies which do not fully lie on the wafer
surface (see figure). These partially
patterned dies don't represent complete
ICs, so they cannot be sold as functional
parts.

Refinements of this simple formula


typically add an edge correction, to
account for partial dies on the edge,
which in general will be more significant
when the area of the die is large
compared to the total area of the wafer.
In the other limiting case (infinitesimally
small dies or infinitely large wafers), the
edge correction is negligible.
The correction factor or correction term
generally takes one of the forms cited by
De Vries:[42]

(area ratio -

circumference/(die diagonal length))


or

(area ratio scaled by an exponential


factor)

or

(area ratio scaled by a polynomial


factor).
Studies comparing these analytical
formulas to brute-force computational
results show that the formulas can be
made more accurate, over practical
ranges of die sizes and aspect ratios, by
adjusting the coefficients of the
corrections to values above or below
unity, and by replacing the linear die
dimension with
(average side length) in the case of dies
with large aspect ratio:[42]

or
or

Crystalline orientation …

Diamond cubic crystal structure of a silicon unit cell

Flats can be used to denote doping and


p g
crystallographic orientation. Red represents material
that has been removed.

Wafers are grown from crystal having a


regular crystal structure, with silicon
having a diamond cubic structure with a
lattice spacing of 5.430710 Å
(0.5430710 nm).[43] When cut into
wafers, the surface is aligned in one of
several relative directions known as
crystal orientations. Orientation is
defined by the Miller index with (100) or
(111) faces being the most common for
silicon.[43] Orientation is important since
many of a single crystal's structural and
electronic properties are highly
anisotropic. Ion implantation depths
depend on the wafer's crystal orientation,
since each direction offers distinct paths
for transport.[44]

Wafer cleavage typically occurs only in a


few well-defined directions. Scoring the
wafer along cleavage planes allows it to
be easily diced into individual chips
("dies") so that the billions of individual
circuit elements on an average wafer can
be separated into many individual
circuits.

Crystallographic orientation
notches

Wafers under 200 mm diameter have
flats cut into one or more sides indicating
the crystallographic planes of the wafer
(usually a {110} face). In earlier-
generation wafers a pair of flats at
different angles additionally conveyed
the doping type (see illustration for
conventions). Wafers of 200 mm
diameter and above use a single small
notch to convey wafer orientation, with
no visual indication of doping type.[45]

Impurity doping …

Silicon wafers are generally not 100%


pure silicon, but are instead formed with
an initial impurity doping concentration
between 1013 and 1016 atoms per cm3 of
boron, phosphorus, arsenic, or antimony
which is added to the melt and defines
the wafer as either bulk n-type or p-
type.[46] However, compared with single-
crystal silicon's atomic density of 5×1022
atoms per cm3, this still gives a purity
greater than 99.9999%. The wafers can
also be initially provided with some
interstitial oxygen concentration. Carbon
and metallic contamination are kept to a
minimum.[47] Transition metals, in
particular, must be kept below parts per
billion concentrations for electronic
applications.[48]

Compound semiconductors
While silicon is the prevalent material for
wafers used in the electronics industry,
other compound III-V or II-VI materials
have also been employed. Gallium
arsenide (GaAs), a III-V semiconductor
produced via the Czochralski process,
Gallium nitride (GaN) and Silicon carbide
(SiC), are also common wafer materials,
with GaN and Sapphire being extensively
used in LED manufacturing.[13]

See also
Die preparation
Epitaxial wafer
Epitaxy
Klaiber's law
Monocrystalline silicon
Polycrystalline silicon
Rapid thermal processing
RCA clean
SEMI font
Silicon on insulator (SOI) wafers
Solar cell
Solar panel
Wafer bonding

References
1. Laplante, Phillip A. (2005). "Wafer" .
Comprehensive Dictionary of Electrical
Engineering (2nd ed.). Boca Raton: CRC
Press. p. 739. ISBN 978-0-8493-3086-5.
2. Reinhard Voelkel (2012). "Wafer-scale
micro-optics fabrication".
doi:10.1515/aot-2012-0013 .
3. T. Doi; I.D. Marinescu; Syuhei Kurokawa
(2012). Advances in CMP Polishing
Technologies, Chapter 6 - Progress of the
Semiconductor and Silicon Industries –
Growing Semiconductor Markets and
Production Areas. Elsevier. pp. 297–304.
doi:10.1016/B978-1-4377-7859-5.00006-
5 .
4. "Martin Atalla in Inventors Hall of Fame,
2009" . Retrieved 21 June 2013.
5. "Dawon Kahng" . National Inventors Hall
of Fame. Retrieved 27 June 2019.
. Lojek, Bo (2007). History of
Semiconductor Engineering. Springer
Science & Business Media. pp. 321–3.
ISBN 9783540342588.
7. Lojek, Bo (2007). History of
Semiconductor Engineering. Springer
Science & Business Media. p. 120.
ISBN 9783540342588.
. "High capacity epitaxial apparatus and
method" . google.com.
9. https://www.tel.com/museum/magazine/
material/150430_report04_03/02.html
10. "Semi" SemiSource 2006: A supplement
to Semiconductor International.
December 2005. Reference Section: How
to Make a Chip. Adapted from Design
News. Reed Electronics Group.
11. SemiSource 2006: A supplement to
Semiconductor International. December
2005. Reference Section: How to Make a
Chip. Adapted from Design News. Reed
Electronics Group.
12. Levy, Roland Albert (1989).
Microelectronic Materials and
Processes . pp. 1–2. ISBN 978-0-7923-
0154-7. Retrieved 2008-02-23.
13. Grovenor, C. (1989). Microelectronic
Materials . CRC Press. pp. 113–123.
ISBN 978-0-85274-270-9. Retrieved
2008-02-25.
14. Nishi, Yoshio (2000). Handbook of
Semiconductor Manufacturing
Technology . CRC Press. pp. 67–71.
ISBN 978-0-8247-8783-7. Retrieved
2008-02-25.
15. "Silicon Solar Cell Parameters" . Retrieved
2019-06-27.
1 . "Evolution of the Silicon Wafer" . F450C.
17. "Archived copy" . Archived from the
original on 2009-02-04. Retrieved
2008-11-26.
1 . "Evolution Of Silicon Wafer | F450C" .
F450C. Retrieved 2015-12-17.
19. "Silicon Wafer" . Archived from the
original on 2008-02-20. Retrieved
2008-02-23.
20. "Intel, Samsung, TSMC reach agreement
about 450mm tech" . intel.com.
21. Presentations/PDF/FEP.pdf ITRS
Presentation (PDF)
22. "450 mm Wafer Handling Systems" .
December 7, 2013. Archived from the
original on December 7, 2013.
23. LaPedus, Mark. "Industry agrees on first
450-mm wafer standard" . EETimes.
24. "The Evolution of AMHS" .
www.daifuku.com.
25. Undeveloped. "semiconductor.net -
Domain Name For Sale" . Undeveloped.
2 . https://www.eetimes.com/collaborative-
advantage-design-impact-of-450mm-
transition/
27. "Lithoguru | Musings of a Gentleman
Scientist" . life.lithoguru.com. Retrieved
2018-01-04.
2 . "Nikon appointing head of precision
equipment business as new president"
(Press release). Japan: Nikon Corp.
semiconportal. 2014-05-20. "Nikon plans
to introduce 450mm wafer lithography
systems for volume production in 2017."
29. LaPedus, Mark (2013-09-13). "Litho
Roadmap Remains Cloudy" .
semiengineering.com. Sperling Media
Group LLC. Retrieved 2014-07-14. "Nikon
planned to ship 'early learning tools' by
2015. 'As we have said, we will be
shipping to meet customer orders in
2015,' said Hamid Zarringhalam, executive
vice president at Nikon Precision."
30. "ASML 2013 Annual Report Form (20-F)"
(XBRL). United States Securities and
Exchange Commission. February 11,
2014. "In November 2013, following our
customers’ decision, ASML decided to
pause the development of 450 mm
lithography systems until customer
demand and the timing related to such
demand is clear."
31. https://www.eetimes.com/first-450-mm-
fabs-to-ramp-in-2017-says-analyst/
32. https://www.eetimes.com/construction-
of-450mm-fab-well-underway/
33. "450mm May Never Happen, says Micron
CEO" . electronicsweekly.com. 11
February 2014.
34. "Intel says 450 mm will deploy later in
decade" . 2014-03-18. Retrieved
2014-05-31.
35. LaPedus, Mark (2014-05-15). "Is 450mm
Dead In The Water?" .
semiengineering.com. California: Sperling
Media Group LLC. Archived from the
original on 2014-06-05. Retrieved
2014-06-04. "Intel and the rest of the
industry have delayed the shift to 450 mm
fabs for the foreseeable future, leaving
many to ponder the following question—Is
450 mm technology dead in the water?
The answer: 450 mm is currently treading
water."
3 . "MW 300GT | Wafer Cases | Shin-Etsu
Polymer Co., Ltd" . www.shinpoly.co.jp.
37. "SMIF Pod-Chung King Enterprise Co.,
Ltd" . www.ckplas.com.
3 . "Wafer Cassette-Chung King Enterprise
Co., Ltd" . www.ckplas.com.
39. "Standing out from the Crowd on 450mm |
450mm News and Analysis" .
40. "H-Square Ergolift Cleanroom Lift Carts" .
www.h-square.com. Archived from the
original on 2019-05-27. Retrieved
2019-05-27.
41. Undeveloped. "semiconductor.net -
Domain Name For Sale" . Undeveloped.
Archived from the original on 2018-08-
21. Retrieved 2018-08-20.
42. Dirk K. de Vries (2005). "Investigation of
gross die per wafer formulas". IEEE
Transactions on Semiconductor
Manufacturing. 18 (February 2005): 136–
139. doi:10.1109/TSM.2004.836656 .
S2CID 32016975 .
43. O'Mara, William C. (1990). Handbook of
Semiconductor Silicon Technology .
William Andrew Inc. pp. 349–352.
ISBN 978-0-8155-1237-0. Retrieved
2008-02-24.
44. Nishi, Yoshio (2000). Handbook of
Semiconductor Manufacturing
Technology . CRC Press. pp. 108–109.
ISBN 978-0-8247-8783-7. Retrieved
2008-02-25.
45. "Wafer Flats" . Retrieved 2008-02-23.
4 . Widmann, Dietrich (2000). Technology of
Integrated Circuits . Springer. p. 39.
ISBN 978-3-540-66199-3. Retrieved
2008-02-24.
47. Levy, Roland Albert (1989).
Microelectronic Materials and
Processes . pp. 6–7, 13. ISBN 978-0-
7923-0154-7. Retrieved 2008-02-23.
4 . Rockett, Angus (2008). The Materials
Science of Semiconductors. p. 13.
ISBN 978-0-387-25653-5.

External links

Wikimedia Commons has media


related to Wafers.
Evolution of the Silicon Wafer by
F450C -An infographic about the
history of the silicon wafer.

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Wafer_(electronics)&oldid=1012390937"

Last edited 4 days ago by 129.127.226.10

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like