Anda di halaman 1dari 9

Cognitive Radios for Spectrum Sharing Anant Sahai, Shridhar Mubaraq Mishra, Rahul Tandra, and Kristen Ann

Woyach Wireless systems require spectrum to operate, but interference is likely if radios in physical proximity simultaneously operate on the same band. Therefore, spectrum is a potentially scarce resource; across the planet today, spectrum is regulated so that most bands are allocated exclusively to a single system licensed to use that band in any given location. However, such static spectrum allocation policies lead to signicant underuse of spectrum [1]. This can be viewed as a kind of regulatory overhead that is paid to get reliable operation. With frequency-agile radios becoming commercially feasible within the next 5-10 years, Cognitive Radio is about making such radios smart enough to share spectrum and reduce the regulatory overhead. This is an impending wireless revolution that draws upon many signal-processing areas including robust detection, sensor networks, as well as the design of incentives and waveforms. This is a short column touching on the issues; further technical details/references can be found in [2]. T HE OPPORTUNITY
IN THE TELEVISION BANDS

Right now, there is signicant excitement surrounding the broadcast television bands. The Federal Communications Commission (FCC) has started considering dynamic approaches for spectrum sharing and the IEEE has launched the 802.22 standards process to use TV-band spectrum holes for enabling wide-area Internet service [3], [4]. This context is illustrated in Figure 1. The background of Figure 1 is a map of the continental USA with the shading representing the population density. The red dots indicate the locations of all TV transmitters while the purple dots correspond to transmitters for channel 40. The green zone on the left zooms in on the San Francisco Bay Area to show the footprints where different stations can be received with an electric eld strength above 41.19dBu for 50% of the locations more than 90% of the time. From this picture, it is clear that spectrum holes are inevitable. Just as a vase can be lled with rocks and still have plenty of room for sand, there is always going to be room for non-interfering radio transmissions in the interstices between channel footprints [5]. The little dark circle represents the interference footprint for channel 40 (where the interference could exceed 2.5 times the thermal noise level of -106dBm more than 10% of the time for more than 50% of the locations) of a hypothetical 802.22 base-station transmitting at 4W from a height of 75m. Just below, a real spectrum scan is shown taken by our group in Berkeley. The local channels are clearly visible. The plot along the top of Figure 1 shows the number of free television channels on a simulated drive from Berkeley, CA to Washington, DC along Interstate 80. The upper blue curve is the size of the opportunity based on International Telecommunications Union (ITU) models for wireless signal propagation run on data from the FCCs database. The lower tan curve illustrates the challenge in using cognitive radios for spectrum sharing. The tan curve predicts the opportunities that would be identied using the current IEEE

Number of Free Channels

60 40 20 0
Recovered by -116 rule

Actually available

Recovering white space under different rules


1 0.8
Actually available by Area Actually available by Population

CCDF

0.6 0.4

-70

-116 rule by Area

95km

0.2 0

-116 rule by Population

Number of channels recovered Distribution of nearest TV tower


1

20

40

60

200km

Power [dB/bin]

50

CDF

Sampling by Area Sampling by Population

0.5

30 10 -10

600

650

Frequency [MHz]

700

750

800

00

Distance [km]

200

400

600

Figure 1.

The nature of spectrum holes in the television bands. (Sources: the FCC TV database for the

latitude/longitude/elevation/power of TV transmitters, the Global Land One-km Base Elevation database from the National Geophysical Data Center for the average terrain elevation (HAAT) value around each transmitter, ITU-R Rec. P.1546-1 for the propagation models, and the 2000 USA Census for the population gures per zip code and the polygonal models for each zip code).

802.22 approach of having a single cognitive radio take a channel measurement and use the channel only if it is sufciently empty. The current IEEE 802.22 rule requires a sensitivity of -116 dBm. While this might prevent interference to television receivers from unfortunately faded cognitive radios, it does so by imposing a tremendous overhead. In most locations, channels that are actually safe to use will still be above -116 dBm for the majority of cognitive radios that are not experiencing unfortunate fading. A statistical nation-wide perspective is given by the plot overlaid on the Midwest. Sampling the USA uniformly by area, on average 56% of the 67 television channels are free while 22% can be recovered by the -116 dBm rule (the area recovered by the -116 dBm rule was calculated using the ITU F(50, 50)

Time Overhead
Spatial Sensing Overhead (1-WPAR)
Energy Detector 12 SNR walls with noise uncertainty = 1 dB

0.8 0.6
P

Prob. of Finding a Hole

Spatial Overhead
1 0.8 0.6 0.4 0.2 0

~ = exp{- (r - rn)} w(r)


Distance from TV transmitter (km)
200 300 400 450

log10N

150

0.4 0.2 0

With Uncertainty = 1 dB = 0.015 km rn = 157 km N= 0 0.2


-1

4 P = P FA= 0.01 MD

Coherent Detector Pilot Power = 10% Coherence Time = 100 -50 -43.3 -40 SNR walls with noise uncertainty = 0.001 dB -33.3 -30

Without Uncertainty

SNR [dB]
1

-20

-13.3 -10

-3.3

ROC curves below SNR wall


SNR = -6 dB 1

ROC curves above SNR wall


1 0.8 SNR = -2.2 dB

Fear of Harmful Interference (F ) HI

0.4

0.6

0.8

1.0

Quantiles

0.4 0.2

N = 100 N = 200

0.4 0.2 0

N = 100 N = 200

0 -5

Support of Y

5 0

Quantiles

H0

0.8

H1

0.6

N = 25 N = 50 N = 75

0.6
P MD

P MD

N = 25 N = 50 N = 75

H0

H1

0 -5 0 0.2 0.4

0.2

0.4

P FA

0.6

0.8

P FA

0.6

0.8

Support of Y

Figure 2.

Uncertainty leads to limits on robust spectrum sensing and overhead in both time and space.

The dotted lines are without noise uncertainty and the solid ones correspond to what actually happens with noise uncertainty.

propagation model). If the population is sampled instead, the average proportion of free channels drops to 33% but the -116 dBm rule can recover only 10%. The plot overlaid on the Deep South shows why sampling by population makes such a difference: television towers are located near population centers. ROBUST SIGNAL PROCESSING AT THE SPECTRUM
SENSORS : TIME AND SPACE

In a single-radio approach to sensing, even weak television signals must be detected to avoid causing interference because the cognitive radio might just be experiencing an unfortunate fade while its own transmissions would interfere with nearby television receivers that are not faded. The traditional signalprocessing approach is to treat this as a hypothesis-testing problem and to compute a test-statistic. By increasing the amount of time N for which the test-statistic is averaged, the hypotheses can traditionally be distinguished arbitrarily well. However the problem in spectrum sensing is that the two hypotheses are themselves uncertain since we cannot completely trust probabilistic models for the noise. This imposes a limit called the SNR Wall on the sensitivity beyond which a detector cannot function reliably. As the signal to noise ratio (SNR) decreases, the distributional uncertainty imposes additional time-overhead that goes to innity at the wall itself. The cause of this can be seen in Figure 2 by examining the receiver operating characteristic (ROC)

curves in the center. Reliable sensing is impossible below the SNR Wall since, as shown to the left of the ROC curves, the two hypothesized sets of distributions for the observation Y overlap. There is also a spatial component to the sensing overhead. To understand this, a simplied model is constructed that has just a single television station, but uses a weighting function w(r) to capture the probability that a point at distance r from this station belongs to the spectrum hole corresponding to this station. The farther away we go, the more likely it is that we are in the service area of another station (and the band is thus unsafe to use). Let rn be the no-talk radius around the television station (the sum of the protected radius shown in Figure 1 by the big television reception circles and the smaller interference footprint of the cognitive radios themselves). A simple two-parameter exponential model wa (r) = aw(r) = a exp((r rn )) can be t to the empirical amount of the overlap (about 10%) between the no-talk regions corresponding to different stations on channel 38 as well as the total fraction of free bands (55%) in channel 38. This wa can be normalized to w and then sensing algorithms can be evaluated using the simple metric

W P AR =
rn

PF H (r)w(r) rdr

where WPAR stands for the weighted probability of area recovered and PF H (r) is the probability that a given spectrum-sensing rule nds an opportunity at a distance r from an isolated television station. The spatial overhead of a sensing algorithm is thus measured by 1 W P AR. This calculation is illustrated in the top-right corner of Figure 2 and we can see that this spatial overhead has a natural tradeoff with the fear (denoted by FHI ) of the wireless fading uncertainty causing harmful interference to the protected television receivers. For example, an FHI of 0.01 means that we must avoid causing interference except in the 1% worst fading events. The -116 dBm rule corresponds to an FHI 2 104 . The SNR Wall phenomenon makes the spatial overhead go to one whenever the FHI is too low. But even ideal single-user sensing has a large spatial overhead at low values of FHI . W HY WE NEED
SPECTRUM SENSING NETWORKS : THE POWER OF COOPERATION

As predicted, the -116 dBm rule of the IEEE 802.22 standard recovers little open spectrum because it is based on single-user single-band sensing and must budget for rare fades. The way around this problem is to exploit the diversity that exists across different radios. Any individual radio might be deeply faded, but it seems unlikely that all cognitive radios in the vicinity will simultaneously be deeply faded. The power of cooperative sensing is shown in the rst two plots of Figure 3. Cooperative rules can recover a lot more area for any given channel and hence more channels at any given location. Performance improves as the number M of independently-faded cooperating radios increases. The Achilles heel of single-band cooperation is shown in the rightmost plot of Figure 3. Fading that might be correlated across users signicantly increases the spatial overhead. The possibility that all sensors

Cooperation
Spatial Sensing Overhead (1- WPAR) Spatial Sensing Overhead (1- WPAR)
1.0 0.8 0.6 0.4 0.2 0.0 10
-4

Spatial Sensing Overhead (1- WPAR)

Scaling
1.0 0.8 0.6 0.4 0.2 0.0

Correlation Uncertainty
1.0 0.8 0.6 0.4 0.2 0.0

F = 0.01 HI

M = 10

M=1 M=2 M=5 M = 10

Empirical performance under - 116 dBm rule (channel 38)

mean= - 70 dBm, std. dev =1

OR rule ML rule

0.8 Correlation uncertainty 0.5 0.2 0

10

-3

10

-2

10

-1

10

10

10

10

10

10

-4

10

-3

10

-2

10

-1

10

mean= -120 dBm, std. dev =2.5

Fear of Harmful Interference (FHI)

Number of Cooperating Users (M)

Fear of Harmful Interference (FHI)

Figure 3.

Understanding the promise/pitfalls of cooperative spectrum sensing. The OR rule declares the

channel to be occupied whenever any of the radios declares the primary to be present. The OR rule only requires limited information about the fading distribution. The Maximum Likelihood (ML) rule uses the average signal power across different sensors as its test statistic and hence requires complete knowledge of the fading distribution [5].

are simultaneously faded cannot be ruled out by mere averaging across sensors. While wireless multipath fading is largely independent for physical reasons, shadowing can be correlated across radios. For example, everyone might go inside when it rains. At rst glance, this appears to be insurmountable. However, the cartoon at the left of Figure 3 illustrates a key insight. While shadowing may be correlated across radios, it is also correlated across frequencies for a single radio! For example, an indoor user will be shadowed relative to television stations and GPS satellites. By exploiting this correlation, multiband sensing can identify and combine sensing information only from those users who are not experiencing severe shadowing. This has the potential to largely eliminate the fear of correlated fading and the resulting spatial overhead [5]. I NCENTIVES AND REGULATION For cognitive radios to move out of the lab, there must be a way to certify the radios and have assurance that they will behave well in the eld. The challenge here is to decide what to certify. For single-user sensing, one could imagine certifying a cognitive radio if it has the appropriate sensitivity and only uses the band when the sensor approves. But certifying the correctness of an implementation of a dynamic protocol that nds neighbors and cooperates with them in the eld seems very difcult. An alternative is to move towards light-handed regulations with minimalist certication and let natural incentives dictate that rational users will not want to cause harmful interference. Figure 4 shows an approach in which cognitive techniques are viewed as bandwidth ampliers that allow a radio to stake its own home band in order to potentially gain access to many other empty bands. A radio is just certied to obey

Global Jail, Util./step = 0 Home Band, Util./step = Cog. Band 1, Util./step = 1 Cog. Band 2, Util./step = 1

TX No Cheat
Home and Two Cog. Bands available Utility = + 2 No Cog. Bands free. Use only Home Utility = False alarm on Band 2. Use only Home Utility = Cheat on unavailable Cog. Band Utility = + 2 In jail. No use of Home or Cog. Bands Utility = 0 In jail Utility = 0 In jail Utility = 0 Out of jail, no Cog. Bands available Utility = One Cog. Band available Utility = + 1

qN q

No TX False Alarm p
1

Ptx = q/(q+p)

Band 3 Band 2 Band 1


Primary
Secondary

Ppen

pN

Band B
Primary

Ppen to incentivize no cheating


Pcatch = 1 Pcatch = 0.5 Pcatch = 0.1

0.5

B=3
0 0.1 0.2

Pwrong 0.3

0.4

0.5

Ppen to incentivize no cheating


Pcatch = 1 Pcatch = 0.5 Pcatch = 0.1

Cheat

TX Legal TX No Cheat

q1

No TX

Ppen
0.5

False Alarm Legal TX

Pwrong = 0.03
1 2 4

Expansion

10

Cognitive

Pc at c
ng Pwro

Utility of the cognitive user


3 2 1 0 0 10 1

t ch

Ppen

Global Jail
Overhead cost of bandwidth expansion
Pwrong = 0.001 Pwrong = 0.005
16 20

Pca

Pw g ron

Cheat

Utility
0.5

Fraction of time in jail

Ptx = 0.55 Pcatch = 1 Pwrong = 0.03 0


30

Expansion

20

Time

40

Maximal bandwidth expansion


Pcatch = 1

30

Pwrong = 0.01

Expansion

Pwrong = .02
20

Expansion

12

Ptx = 0.9 Ptx = 0.55 Ptx = 0.1 Pwrong scales with expansion Pcatch = 1

Pcatch = 1 Pcatch = 0.5

=1

Pwrong = 0.035
10

Pwrong = 0.06 Pwrong = 0.1


0 0.1 0.2 0.3 0.4

Maximal Expansion Ptx = 0.55 Pcatch = 1


0.5

Pcatch = 0.1

Avg Use without Cog. user = 4/9 Avg Use with Cog. user = 6/9 Avg Utility for Cog. user = (6+5)/9

Ptx = 0.55
0 0.1 0.2

Overhead

Pwrong

0.3

0.4

0.5

Figure 4.

Cognitive radios for bandwidth expansion by selsh users.

a wireless command to go to jail for a period of time during which it loses access to all bands, including its own home band. This command is issued when the radio is caught cheating (causing interference). The fear of prison must be high enough to keep the selsh radios honest [6]. On the left-hand side of Figure 4, a timeline is shown in which a cognitive radio is caught and sent to jail. In the top left, a Markov chain is shown for modeling the behavior of the licensed users in different bands as well as the cognitive radios choice to cheat or not to cheat. P pen controls how long the jail sentences are. The top right of Figure 4 shows how the sentences must get harsher as either the temptation (number of bands B) increases or as the probability P wrong of wrongful conviction increases. Once P pen is set, the cognitive user can calculate its expected utility from an expansion factor of
B .

It is not worth

expanding beyond a certain point since the utility gained from additional bands would be offset by the increasing time spent in jail due to the few inevitable wrongful convictions. The bottom right corner of Figure 4 shows the maximal bandwidth expansion as a function of P wrong and the probability P catch of being caught when truly cheating. However, there is an overhead due to

Network ID User ID Device ID TX Identity: Band 1


Cannot transmit

...

TX Identity: Band 2 TX Identity: Band 3

Percentage increase in Primary errors

1000

Minimum enforcement overhead

Enforcement overhead

800

5% background error in Primary link Pcatch = 0.9 Pwrong = 0.005

0.5

0.5

Time steps until conviction = 3000

0.4 50% 0.3 65% 0.2 100% 0.1 200% increase in primary errors 0 2000 4000 6000 8000 10000

0.4

600

0.3 Catch coalition of 4 0.2 Catch coalition of 3 0.1 Catch coalition of 2 0

400

Overhead = 5% Overhead = 10%

200

Overhead = 25%

2000

4000

6000

8000

10000

10 2

10 3

Time steps until conviction

Time steps until conviction

10 5 10 4 Number of users

10 6

10 7

Figure 5.

Identity ngerprints for cognitive radios.

users being wrongfully convicted and thereby being unable to use either their own bands or true spectrum holes. The tradeoff between this overhead and bandwidth expansion is shown in the bottom left of Figure 4. For example, a potential expansion into all 67 of the 6 MHz TV bands by a user staking a single large WiMAX channel of 20 MHz requires a bandwidth expansion of about 20. To keep the wrongful conviction overhead below 10%, Figure 4 reveals that P wrong needs to be about 1% if P catch = 1. At a more realistic P catch of 0.9, the required P wrong must be a very stringent 0.5%. This leads us directly to the second regulatory requirement: a way to reliably identify the source of harmful interference. This was described vividly by Faulhaber as the problem of hit and run radios that he feared would not only preclude the potential commercial impact of cognitive radios, but also rule out any approach that involved a real-time market for wireless spectrum [7]. How can a toll road be sustained without any toll booths or controlled on-ramps? The answer is clear: whether it is a public highway or a toll road, we need license plates to balance the freedom of drivers with the requirements of the community. Wireless identity certication involves the design of the radios waveforms so that appropriate signal processing can recover their identity. The most straightforward approach would be to require the broadcast of an explicit identity beacon. However, this would require the government to mandate a single beacon waveform to be broadcast by all cognitive radios, regardless of their own native waveforms. Not only would this be an added expense, it would also stop certain socially desirable approaches from working at all. For example, radios that tried to use beamforming to avoid causing interference would have their hopes dashed by the interference caused by their government-mandated omnidirectional beacons. Figure 5 shows another approach. Each radio has a unique ngerprint of time-slots that it is not allowed

to use in each band. As shown in the top of Figure 5, this identity code might be a composite of many different aspects (e.g. the network, the human user, the physical device, etc.) of the identity, but it has the property that any radio causing harmful interference will leave its ngerprints behind in the pattern of interference itself. This code can easily be certied in the hardware without constraining the detailed waveforms at the packet level. The overhead imposed by the code is the proportion of slots that must be left silent because during this time, the user is blocked from exploiting a spectrum opportunity [8]. The two bottom left plots in Figure 5 illustrate the tradeoffs between the time to catch a cheater and the level of interference that the licensed users want to guard against. It is easy to catch systems that cause a lot of interference. But if the level of interference is low, convicting a suspect is hard unless we are willing to tolerate a lot of overhead. The bottom right plot in Figure 5 shows information-theoretic lower bounds on the overhead required if the time is constrained to 3000 slots (half a minute if each slot is ten milliseconds long). The overhead increases with the number of identities as well as with the size of the coalitions of simultaneous cheaters. Being able to convict more than one cheater is important to deter the wireless equivalent of looting wherein one cheater will induce everyone else to cheat as well. C ONCLUSIONS The signal processing issues involved in cognitive radios are quite diverse and have led us on a gurative journey from Berkeley, CA to Washington DC. A holistic SP perspective shows that while the goal of reducing the regulatory overhead is admirable, everything will have to be put together in a balanced way in order to realize the true potential of this wireless revolution. ACKNOWLEDGEMENTS We thank the National Science Foundation (grants ANI-326503, CNS-403427, CCF-729122 as well as a Graduate Research Fellowship), Sumitomo Electric and the support of the Center for Circuit & System Solutions (C2S2) Focus Center, one of ve research centers funded under the Focus Center Research Program, a Semiconductor Research Corporation program.. AUTHORS Prof. Anant Sahai (sahai@eecs.berkeley.edu) and his students Mubaraq Mishra (smm@eecs.berkeley.edu), Rahul Tandra (tandra@eecs.berkeley.edu), and Kristen Woyach (kwoyach@eecs.berkeley.edu) are all with the EECS Department at UC Berkeley. R EFERENCES [1] Spectrum policy task force report, Federal Communications Commission, no. 02-135, Nov. 2002. [2] A. Sahai, S. M. Mishra, R. Tandra, and K. A. Woyach, Extended Edition: Cognitive radios for spectrum sharing, Tech Report in preparation, 2008.

[3] Unlicensed Operation in the TV Broadcast Bands, Federal Communications Commission, First Report and Order and Further Notice of Proposed Rulemaking. 06-156, Oct. 2006. [4] C. R. Stevenson, C. Cordeiro, E. Sofer, and G. Chouinard, Functional requirements for the IEEE 802.22 WRAN standard, Tech. Rep., September 2005. [5] R. Tandra, S. M. Mishra, and A. Sahai, What is a spectrum hole and what does it take to recognize one? To appear in the Proceedings of the IEEE, Jan 2009. [6] K. A. Woyach, Crime and punishment for cognitive radios, Masters thesis, UC Berkeley, 2008. [7] G. R. Faulhaber, The future of wireless telecommunications: spectrum as a critical resource, Information Economics and Policy, vol 18, no. 3, pp 256-271, Sep. 2006. [8] G. Atia, A. Sahai, and V. Saligrama, Spectrum Enforcement and Liability Assignment in Cognitive Radio Systems, Proceedings of the 3rd IEEE International Symposium on New Frontiers in Dynamic Spectrum Access Networks, Chicago IL, Oct. 2008.

Anda mungkin juga menyukai