Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

PAPR analysis for OFDM visible light communication

Open Access Open Access

Abstract

Orthogonal frequency-division multiplexing (OFDM) is a practical technology in visible light communication (VLC) for high-speed transmissions. However, one of its operational limitations is the peak-to-average power ratio (PAPR) of the transmitted signal. In this paper, we analyze the PAPR distributions of four VLC OFDM schemes, namely DC-biased optical OFDM (DCO-OFDM), asymmetrically clipped optical OFDM (ACO-OFDM), pulse amplitude modulated discrete multitone (PAM-DMT), and Flip-OFDM. Both lower and upper clippings are considered. We analytically derive the complementary cumulative distribution functions (CCDFs) of the PAPRs of the clipped VLC OFDM signals, and investigate the impact of lower and upper clippings on PAPR distributions. Our analytical results, as verified by numerical simulations, provide useful insights and guidelines for VLC OFDM system designs.

© 2016 Optical Society of America

1. Introduction

Recently, visible light communication (VLC) has emerged as a promising technology for providing indoor wireless services [1]. Utilizing light-emitting diodes (LEDs) [2], VLC can support illumination and communication simultaneously and offers the benefits of high rates, low cost, safety, security, and no electronic interference. Therefore, VLC has been regarded as an important supplement to, or replacement of, traditional RF communication in indoor environments [3–5].

To fulfill high-speed transmissions, orthogonal frequency-division multiplexing (OFDM), which has been widely used in RF communication systems (e.g., LTE), has also been considered in VLC [6]. However, unlike RF communication that can coherently transmit bipolar and complex-valued signals, VLC generally exploits intensity modulation (IM) and direct detection (DD) to generate real and nonnegative LED-driving signals. Hence, traditional RF OFDM techniques cannot be directly applied to VLC and must be specially modified This reason has led to the rise of various VLC OFDM schemes such as DC-biased optical OFDM (DCO-OFDM) [6], asymmetrically clipped optical OFDM (ACO-OFDM) [7], pulse amplitude modulated discrete multitone (PAM-DMT) [8], and Flip-OFDM [9].

A well known problem of OFDM is its high peak-to-average power ratio (PAPR) [10]. To prevent intermodulation among subcarriers and out-of-band radiation, power amplification at the transmitter should operate in its linear region. A high PAPR requires the power amplifier to possess a wide linear region using a large power back-off, which substantially compromises efficiency. Hence, the PAPR issue has received much attention in RF communication. The PAPR distribution of complex-valued baseband RF-OFDM signals has been studied [11] and various PAPR reduction techniques have also been proposed [10,12–14]. On the other hand, only a few works exist that tackle the PAPR problem in VLC OFDM [13–16]. Thus far, there does not exist a comprehensive study of PAPR distributions of common VLC OFDM schemes.

It is worth noting that existing PAPR analysis for RF-OFDM, e.g., [2], does not apply to VLC OFDM. Time-domain VLC OFDM signals must be real and nonnegative. The nonnegativeness is insured by clipping the signal from below (a DC offset may be added, e.g., in DCO-OFDM). Meanwhile, an LED usually imposes a limit on the maximum magnitude of the input signal, in response to eye safety or device considerations [17,18]. Thus, time-domain VLC OFDM signals may also be clipped from above. Such lower and upper clippings shall be taken into PAPR analysis. In the literature, the cumulative distribution function (CDF) of the unclipped DCO-OFDM signal was investigated in [15]. In [16], the distributions of upper and lower PAPRs of DCO-OFDM, instead of the exact PAPR, were derived and the clipping was replaced by a scaling operation. Therefore, the distributions of the PAPRs of the clipped VLC OFDM signals remain unknown.

In this paper, we investigate the PAPR distributions of four VLC OFDM schemes, including DCO-OFDM, ACO-OFDM, PAM-DMT, and Flip-OFDM. Both lower and upper clippings are considered. We derive analytical expressions of the complementary cumulative distribution functions (CCDFs) of the PAPRs of the clipped VLC OFDM signals. We show that the PAPR CCDF is a piecewise function, which is divided by two points for the asymmetrically clipped DCO-OFDM signal and divided by one point for the symmetrically clipped DCO-OFDM signal as well as the clipped ACO-OFDM, PAM-DMT, and Flip-OFDM signals. The (maximum) dividing point is the PAPR ceiling that defines the maximum value of all possible PAPRs. Such a ceiling exists if and only if both lower and upper clippings are performed. We further investigate the impact of lower and upper clippings on PAPRs and show that both the PAPR CCDF and the PAPR ceiling are monotonic with respect to lower and upper clipping bounds, respectively. These results are particularly useful to determine when lower and upper clippings decrease or increase the PAPR. Finally, the derived analytical PAPR CCDFs and the monotonic properties in terms of clipping bounds of the four VLC OFDM schemes are verified by numerical simulations.

Note that clippings not only affect the PAPR but also introduce distortion to the transmitted signal. In this paper, we focus on analyzing the effect of upper and lower clippings on the PAPR with given clipping bounds, signal power, and DC offset. The interested reader is referred to [17–20] for optimization of the clipping, signal power and DC offset.

2. System models

In the section, we introduce the system models of DCO-OFDM, ACO-OFDM, PAM-DMT, and Flip-OFDM.

2.1. DCO-OFDM

Figure 1 shows the transmitter of the DCO-OFDM system [6]. A serial bit stream is split by a serial-to-parallel (S/P) converter and then mapped to quadrature amplitude modulated (QAM) symbols. Let N be the number of subcarriers and Sk be the symbol on the kth subcarrier. To obtain a real time-domain signal, the frequency-domain Hermitian symmetry is required:

Sk=SNk*,k=1,,N21
and the 0th and N2th subcarriers are null, i.e., S0 = SN/2 = 0. Hence, the frame structure of DCO-OFDM is as follows
S=[0,S1,S2,,SN21,0,SN21*,,S2*,S1*].
The time-domain signal is obtained by applying the inverse fast Fourier transform (IFFT) to Sk
sn=1Nk=0N1Skej2πknN=2Nk=1N21Re{Skej2πknN}
for n = 0, . . . , N − 1, where Re{·} denotes the real part of a complex value.

 figure: Fig. 1

Fig. 1 Transmitter block diagram of the DCO-OFDM VLC System.

Download Full Size | PDF

As VLC exploits intensity modulation (IM) to drive an LED, the transmitted signal sn is required to be not only real but also nonnegative. Thus, a lower clipping and a DC bias are necessary. Moreover, an LED may have a limit on the maximum value of the input signal, e.g., for brightness control or fitting in the linear range of the transfer characteristic, so an upper clipping may also be needed. The clipping process can be expressed as

sclip,n=Clip[sn]={BU,snBUsn,BL<sn<BUBL,snBL
where BU and −BL represent the upper and lower clipping bounds. The smaller BU and BL are, the more the signal is clipped. The clipped signal sclip,n is fed into the the digital-to-analog convertor (DAC) resulting in the continuous time signal sclip(t). A DC bias BL is added onto sclip(t) to obtain a nonnegative signal sdc(t) = sclip(t) + BL, which is then used to drive an LED.

The receiver uses a photo detector to convert the transmitted signal, which passes through the optical channel, to an electrical signal and then performs direct detection (DD) to restore the transmitted symbols.

2.2. ACO-OFDM

Figure 2 shows the transmitter of the ACO-OFDM system [7]. In ACO-OFDM, only the odd subcarriers carry data symbols and the even subcarriers are set to zero, i.e., Sk = 0 for k = 0, 2, 4, ..., N − 2. At the same time, ACO-OFDM also satisfies the Hermitian symmetry in the frequency domain and thus has the following frame structure

S=[0,S1,0,,SN21,0,SN21*,,0,S1*].
Therefore, the spectral efficiency of ACO-OFDM is a half of that of DCO-OFDM. After the IFFT, the time-domain signal is obtained for n = 0, . . . , N − 1
sn=1Nk=0N1Skej2πknN=2Nk=1N21Re{Skej2πknN}=2Nt=1N/4Re{S2t1ej2π(2t1)nN}.
Note that the time-domain signal of ACO-OFDM has following symmetric property:
sn+N2=2Nt=1N/4Re{S2t1ej2π(2t1)(n+N2)N}=sn,n=0,,N21.
Therefore, one can transmit the positive part of sn and clip its negative part without losing any information. Such a clipping process plus the upper bound limit can be expressed as
sclip,n=Clip[sn]={BU,snBUsn,0<sn<BU0,sn0.
Since sclip,n is nonnegative, there is no need to add a DC bias. The clipped signal is then converted to the continuous signal sclip(t) for transmitting via an LED.

 figure: Fig. 2

Fig. 2 Transmitter block diagram of the ACO-OFDM VLC System.

Download Full Size | PDF

2.3. PAM-DMT

Figure 3 shows the transmitter of the PAM-DMT system [8]. Different from DCO-OFDM and ACO-OFDM that use complex QAM symbols, PAM-DMT uses PAM symbols as the imaginary parts of transmitted symbols. Specifically, the transmitted symbol on subcarrier k is given by Sk = jAk for k=1,,N21, where Ak is a real PAM symbol. To satisfy the Hermitian symmetry in the frequency domain, the PAM-MDT frame has the following structure

S=[0,jA1,jA2,,jAN21,0,jAN21,,jA2,jA1].
Thus, PAM-DMT has the same spectral efficiency as ACO-OFDM. After the IFFT, the time-domain signal is obtained
sn=1Nk=0N1Skej2πknN=1Nk=1N21[jAkej2πknNjAkej2π(Nk)nN]=2Nk=1N21Aksin2πknN
for n = 0, . . . , N − 1. Note that the time-domain signal of PAM-DMT is also symmetric
sNn=2Nk=1N21Aksin2πk(Nn)N=snn=1,,N21
and s0 = sN/2 = 0. Therefore, similar to ACO-OFDM, one can transmit the nonnegative part of sn and clip its negative part without adding a DC bias. The clipped signal sclip,n is obtained by using the same clipping function in (5) on sn.

 figure: Fig. 3

Fig. 3 Transmitter block diagram of the PAM-DMT VLC System.

Download Full Size | PDF

2.4. Flip-OFDM

Figure 4 shows the transmitter of the Flip-OFDM system [9]. The transmitting process of Flip-OFDM is the same with that of DCO-OFDM until the IFFT, thus resulting in the same time-domain signal sn as in (2). Then, sn is divided into the positive and negative parts as sn=sn++sn where

sn+={sn,ifsn00,otherwiseandsn={sn,ifsn<00,otherwise.
These two parts are transmitted through two frames with the first frame sn1=sn+ and the second frame sn2=sn. Hence, the spectral efficiency of Flip-OFDM is equal to that of ACO-OFDM or PAM-DMT, and a half of that of DCO-OFDM.

 figure: Fig. 4

Fig. 4 Transmitter block diagram of the Flip-OFDM VLC System.

Download Full Size | PDF

Let s¯n=[sn1,sn+N2] be the expended time-domain signal. Since n is nonnegative, only an upper clipping is needed

sclip,n=Clip[s¯n]={BU,snBUsn,sn<BU
and no DC bias is added. At the receiver, the received time-domain signals of two successive frames, say yn1 and yn2, will be combined as yn=yn1yn2 for demodulation.

3. PAPR analysis

In this section, we analyze the PAPRs of the clipped signals of the above four VLC OFDM schemes. The clipped signal sclip,n will be converted to an analog signal and amplified by a power amplifier to drive an LED (in DCO-OFDM the DC bias is added to the amplified signal). As introduced in Section 1, the implementation of power amplification is significantly influenced by the PAPR of the clipped OFDM signal. According to [10, 11], The PAPR of the clipped signal sclip,n is defined as

PAPR=max0nN1|sclip,n|2E[|sclip,n|2]
which is a random variable and characterized by its CCDF, i.e., the probability P(PAPR > x). In the following, we will derive the PAPR CCDFs of the four VLC OFDM schemes and investigate the impact of clipping.

3.1. PDFs of the clipped signals

To derive the PAPR CCDF, the probability density function (PDF) of the clipped signal is needed. Note that the frequency-domain symbols Sk are in general independently identically distributed (i.i.d.) with zero mean. From the central limit theorem (CLT), for a large subcarrier number N, the time-domain signal sn obeys a zero-mean Gaussian distribution [11, 15–17, 21], i.e., sn~𝒩(0,σs2), with the PDF p(w)=12πσsexp{w22σs2}, where σs2 is the power (variance) of the time-domain signal. This holds for practical subcarrier numbers (typically N ≥ 128) [6].

Define the normalized lower and upper clipping bounds l ∈ (−∞, 0] and u ∈ [0, +∞) as

l=BLσsandu=BUσs.
Denote by g(x)=12πex22 and Φ(x)=x12πet22dt the PDF and CDF of the standard normal distribution, respectively. Let δ(x) be the Dirac delta function and u(x) be the step function. The clipped DCO-OFDM signal sclip,n follows a truncated Gaussian distribution [22], whose PDF is given by
fdco(w)=Φ(l)δ(w+BL)+12πσsew22σs2[u(w+BL)u(wBU)]+[1Φ(u)]δ(wBU).
Note that the clipping process in ACO-OFDM, PAM-DMT, or Flip-OFDM can be regarded as a simplified version in DCO-OFDM by setting BL = 0. Thus, the PDF of the clipped ACO-OFDM, PAM-DMT, or Flip-OFDM signal is obtained by setting l = 0 in fdco(w), leading to
faco(w)=12δ(w)+12πσsew22σs2[u(w)u(wBU)]+[1Φ(u)]δ(wBU).

3.2. Average power of the clipped signal

Given the PDF of the clipped signal, we can compute the average electrical power of sclip,n, which is the denominator of the PAPR. In particular, the average electrical power of the clipped DCO-OFDM signal is

Pdcoele=E[|sclip,n|2]=E[sclip,n2]=+w2fdco(w)dw=σs2c(l,u)
where c(l, u) is defined as
c(l,u)=(l21)Φ(l)(u21)Φ(u)+u2+lg(l)ug(u).
The average electrical power of the clipped ACO-OFDM, PAM-DMT, or Flip-OFDM signal is then obtained by setting l = 0 in (9), leading to Pacoele=σs2c(0,u).

Lemma 1 c(l, u) ∈ [0, 1] is decreasing in l (increasing in BL) and increasing in u (BU).

Proof. It can be verified that ∂c(l, u)/∂l = 2lΦ(l) ≤ 0 and ∂c(l, u)/∂u = 2u(1 − Φ(u)) ≥ 0. Therefore, c(l, u) is decreasing in l and increasing in u, and 0 = c(0, 0) ≤ c(l, u) ≤ c(−∞, +∞) = 1. □

It follows from Lemma 1 that Pdcoele and Pacoele are increasing in BU and Pdcoele is increasing in BL. Given the same upper clipping bound BU, we have PdcoelePacoele. This is consistent with the intuition that the more the signal is clipped, the less the average electrical power is. On the other hand, it is not straightforward to see the impact of clipping on the PAPR, since the peak power is also reduced by upper and lower clippings.

Moreover, we can also compute the average optical power of the clipped signal, which is given by E[sclip,n]. In particular, the average optical power of the clipped DCO-OFDM signal is

Pdcoopt=E[sclip,n]+BL=+wfdco(w)dw=σs[lΦ(l)+g(l)g(u)+u(1Φ(u))]+BL
and the average optical power of the clipped ACO-OFDM, PAM-DMT, or Flip-OFDM signal is given by Pacoopt=σs[12πg(u)+u(1Φ(u))].

3.3. PAPR distribution of DCO-OFDM

Given the PDF and average electrical power of the clipped DCO-OFDM signal, we are able to provide an analytical expression of the PAPR CCDF for DCO-OFDM.

Theorem 1 The PAPR CCDF of the clipped DCO-OFDM signal is

Fdco(x)=P(PAPR>x)={1[2Φ(c(l,u)x)1]N,0x<θmin1[Φ(c(l,u)x)]N,θminx<θmax0,xθmax
where θmin = min{θL, θU} and θmax = max{θL, θU} with
θL=l2c(l,u)andθU=u2c(l,u).
Proof. See Appendix A1. □

From Theorem 1, the PAPR CCDF of the clipped DCO-OFDM signal, i.e., Fdco(x), is a piecewise function divided by θmin and θmax, which are jointly determined by the normalized lower and upper clipping bounds l and u (or BL and BU). If BLBU, then θmin = θL and θmax = θU, otherwise θmin = θU and θmax = θL. Note that Fdco(x) is not continuous at θmin and θmax. In particular, Fdco(x) = 0 if xθmax, i.e., θmax is the ceiling of the PAPR. Such a PAPR ceiling exists if only if both lower and upper clippings are performed. If the upper and lower clippings are symmetric, i.e., −l = u(BL = BU), then the PAPR CCDF of the clipped DCO-OFDM signal simplifies to

Fsym(x)={1,[2Φ(c(u,u)x)1]N,0x<θ0,xθ
where θ = u2/c(−u, u). In this case, there is only one dividing point θ, which is the PAPR ceiling.

Proposition 1 θL is decreasing in l and u, respectively, and θU is increasing in l and u, respectively. For the symmetric clipping (l = u), θ is increasing in u.

Proof. See Appendix A2. □

Proposition 1 indicates that the dividing points θmin and θmax are monotonic in the upper and lower clipping bounds, but depending on the relation between l and u (or BL and BU). Specifically, If −l < u (BL < BU), then θmin = θL is decreasing in l and u and θmax = θU is increasing in l and u, and vice versa. This reveals that in the asymmetric case, a single-side clipping may not reduce but may enlarge the PAPR ceiling. Indeed, for −l < u, θmax = θU is increasing in l for fixed u, i.e., the more the signal is clipped from below, the larger the PAPR ceiling is. Similarly, for −l > u, θmax = θL is decreasing in u for fixed l, i.e., the more the signal is clipped from above, the larger the PAPR ceiling is. Hence, an effective way to reduce the PAPR ceiling θmax is to perform both lower and upper clippings. This can be clearly seen from the symmetric clipping −l = u (BL = BU), where the PAPR ceiling θ is increasing in u, implying that the more the signal is (symmetrically) clipped, the less the maximum PAPR is.

Proposition 2 For 0 ≤ x < θmin or θminx < θmax, Fdco(x) is increasing in l and decreasing in u, respectively. For 0 ≤ x < θ, Fsym(x) is decreasing in u.

Proof. The monotonicity of Fdco(x) in terms of l and u follows directly from Lemma 1. For Fsym(x), it can be verified that c′(−u, u) = 4u(1 − Φ(u)) ≥ 0, so c(−u, u) is increasing in u. Therefore, Fsym(x) is decreasing in u for 0 ≤ x < θ. □

Proposition 2 indicates that the PAPR CCDF of DCO-OFDM is monotonic in the clipping bounds within some range. Specifically, for 0 ≤ x < θmin or θminx < θmax, the smaller the clipping bounds BU and BL are, i.e., the more the DCO-OFDM signal is clipped, the larger the PAPR could be. However, such a monotonicity does not hold over the whole PAPR range. In fact, from Propositions 1 and 2, we can see a tradeoff in the PAPR caused by clipping. Take the symmetric clipping for example, the more the signal is clipped (i.e., u is smaller), the larger the PAPR could be within [0, θ). On the other hand, a smaller u leads to a smaller θ, i.e., a smaller PAPR ceiling.

From Theorem 1, we can readily obtain the PAPR CCDF of the unclipped signal sn. Specifically, by setting BL = BU = +∞ (l = −∞ and u = +∞), which leads to c(−∞, +∞) = 1 and θmin = θmax = +∞, we obtain

Func(x)=1[2Φ(x)1]N,x0.
Proposition 3 Let α = min{θmax, x0} and β = max{θmin, x0}, where x0 is the root of the equation Φ(c(l,u)x)=2Φ(x)1. Then,
Fdco(x){Func(x),x<θminorβx<θmaxFunc(x),xθmaxorθminxα
and
Fsym(x){Func(x),0x<θFunc(x),xθ.
Proof. See Appendix A3. □

Proposition 3 indicates that the PAPR CCDF of the clipped DCO-OFDM signal could be higher or lower than that of the unclipped signal within different PAPR ranges. This is actually consistent with Propositions 1 and 2 in the sense that clipping has both positive and negative impacts on the PAPR. The positive aspect is that (double-side) clipping leads to a PAPR ceiling restricting the maximum PAPR, and the negative aspect is that clipping increases the probability of large PAPRs below the ceiling. Note that the nonlinear equation Φ(c(l,u)x)=2Φ(x)1 has a unique root, since Φ(c(l,u)x) and 2Φ(x)1 are both monotonically increasing in x.

3.4. PAPR distributions of ACO-OFDM and PAM-DMT

Now, we investigate the PAPR distributions of the clipped ACO-OFDM and PAM-DMT signals.

Theorem 2 The PAPR CCDF of the clipped ACO-OFDM signal is

Faco(x)=P(PAPR>x)={1[2Φ(c(0,u)x)1]N2,0x<θ0,U0,xθ0,U
where θ0,U = u2/c(0, u).

Proof. See Appendix A4. □

It is worth pointing out that Faco(x) of ACO-OFDM cannot be obtained from Fdco(x) of DCO-OFDM. Indeed, if one sets l = 0 in Fdco(x), he will obtain Fdco(x)=1[Φ(c(0,u)x)]N for 0 ≤ x < θ0,U, which is different from Faco(x)=1[2Φ(c(0,u)x)1]N2 for 0 ≤ x < θ0,U. The reason is that there is a symmetry in the (unclipped) ACO-OFDM signal sn (see (4)). Therefore, Faco(x) has to be derived in a different way.

Follow the similar steps, we can show that θ0,U is increasing in u, i.e., the more the ACO-OFDM signal is clipped from above, the smaller the PAPR ceiling is. On the other hand, Faco(x) is decreasing in u for 0 ≤ x < θ0,U, i.e., the more the ACO-OFDM signal is clipped, the more likely a large PAPR appears within [0, θ0,U). In the case that there is no upper clipping, i.e., u = +∞, we obtain c(0, +∞) = 1/2 and Faco(x)=1[2Φ(x/2)1]N2 for x ≥ 0.

Theorem 3 The PAPR CCDF of the clipped PAM-DMT signal is

Fpam(x)=P(PAPR>x)={1[2Φ(c(0,u)x1]N21,0x<θ0,U0,xθ0,U
where θ0,U = u2/c(0, u).

The PAPR CCDF of PAM-DMT can be derived in a similar way as that of ACO-OFDM. Indeed, the PAM-DMT signal is clipped in the same manner as ACO-OFDM in (5) and the unclipped PAM-DMT signal also has a symmetry. In particular, the time-domain PAM-DMT signal sn satisfies sn = −sNn for n=1,.N21, which, though a bit different from the symmetry of the ACO-OFDM signal in (4), enables us to use the similar reasoning in deriving the CCDF. The reason why Fpam(x)=1[2Φ(c(0,u)x1]N21 instead of 1[2Φ(c(0,u)x1]N2 for 0 ≤ x < θ0,U is that for PAM-DMT s0 and sN/2 are zero. In practice, for a large number of subcarriers, the PAPR CCDFs of PAM-DMT and ACO-OFDM tend to be equal. Apparently, Fpam(x) shares the same monotonicity in the upper clipping bound u as Faco(x). In the case that there is no upper clipping, Fpam(x) simplifies to Fpam(x)=1[2Φ(x/2)1]N21 for x ≥ 0.

3.5. PAPR distribution of Flip-OFDM

A complete time-domain Flip-OFDM frame consists of two N-point frames, which correspond to the positive and (flipped) negative parts of the time-domain signal sn, respectively. Therefore, the PAPR of the clipped Flip-OFDM signal shall be

PAPR=max0n2N1|sclip,n|2E[|sclip,n|2]
whose CCDF is given in the following result.

Theorem 4 The PAPR CCDF of the clipped Flip-OFDM signal is

Fflip(x)=P(PAPR>x)={1[2Φ(c(0,u)x1)]N,0x<θ0,U0,xθ0,U
where θ0,U = u2/c(0, u).

Proof. See Appendix A5. □

Note that the positive and negative parts of the time-domain signal sn are correlated. This is the reason that the PAPR CCDF of Flip-OFDM has a similar form with those of ACO-OFDM and PAM-DMT, both of which introduce correlation in the time-domain signal as a result of the time-domain symmetry. On the other hand, as a Flip-OFDM frame contains two N-point frames, the exponent in Fflip(x) is N instead of N/2 (as in ACO-OFDM) for 0 ≤ x < θ0,U. Similar to Faco(x) and Fpam(x), Fflip(x) is decreasing in u for 0 ≤ x < θ0,U, while θ0,U is increasing in u. In the case that there is no upper clipping, Fflip(x) simplifies to Fflip(x)=1[2Φ(x/2)1]N for x ≥ 0. From Theorems 2, 3, and 4, we can obtain the following relation between the PAPR CCDFs of ACO-OFDM, PAM-DMT, and Flip-OFDM

Faco(x)Fpam(x)Fflip(x)
since 02Φ(c(0,u)x)11. In practice, Faco(x) ≈ Fpam(x) ≤ Fflip(x).

4. Simulation results

In this section, we verify the derived analytical results through numerical simulations. We first show the values of the three important parameters c(l, u), θL and θU versus the normalized clipping bounds l and u in Fig. 5. Note that c(l, u) is proportional to the average electrical power of the clipped signal (see Section 3.2). From Fig. 5(a), one can observe that c(l, u) is increasing in u and decreasing in l, which is consistent with Lemma 1. θL and θU are the dividing points of the piecewise PAPR CCDF of the clipped DCO-OFDM signal and PAPR ≤ θmax = max{θL, θU}. It can be observed from Fig. 5(b) that θL is decreasing in l and u and θU is increasing in l and u, which complies with Proposition 1.

 figure: Fig. 5

Fig. 5 (a) values of c(l, u) and (b) values of θL and θU for different l and u.

Download Full Size | PDF

Figure 6 displays the PAPR CCDFs of the four VLC OFDM schemes with different subcarrier numbers, where DCO-OFDM, ACO-OFDM, and Flip-OFDM exploit 4-QAM modulation and PAM-DMT exploits 4-PAM modulation. For DCO-OFDM, PAM-DMT, and Flip-OFDM, the analytical and simulated PAPR CCDFs match well even if there are only N = 128 subcarriers. In ACO-OFDM the gap between the analytical and simulated CCDFs is a bit larger than that in DCO-OFDM, PAM-DMT, or Flip-OFDM for N = 128. This is because the number of independent frequency-domain symbols is N/2 in DCO-OFDM, PAM-DMT, and Flip-OFDM but N/4 in ACO-OFDM. Therefore, the Gaussian approximation of the time-domain signal is more accurate in DCO-OFDM, PAM-DMT, and Flip-OFDM than in ACO-OFDM given the same subcarrier number. Indeed, when the subcarrier number is equal to or larger than 256, the analytical and simulated ACO-OFDM PAPR CCDFs also match well. Note that analytical and simulated CCDFs have exactly the same PAPR ceiling, which indicates the accuracy of our analytical results too.

 figure: Fig. 6

Fig. 6 PAPR CCDFs of the four VLC OFDM schemes with different subcarrier numbers.

Download Full Size | PDF

In Fig. 7, we plot the simulated and analytical PAPR CCDFs of the clipped and unclipped DCO-OFDM signals with N = 1024 subcarriers. For the asymmetric clipping, the PAPR CCDF has two dividing points, i.e., θmin and θmax. For l = −5 and u = 0, . . . , 4, θmin = θU is increasing in u and θmax = θL is decreasing in u. For u = 5 and l = 0, . . . , −4, θmin = θL is decreasing in l and θmax = θU is increasing in l. It is consistent with Proposition 1 and also demonstrates that single-side clippings may increase the PAPR ceiling, i.e., the more the signal is clipped from one side (above or below), the larger θmax becomes. For the symmetric clipping, the PAPR CCDF has only one dividing point θ, which is increasing in u. In this case, the more the signal is clipped, the smaller the PAPR ceiling θ is. The PAPR CCDF of the clipped DCO-OFDM signal could be smaller or larger than that of the unclipped signal (l = −∞, u = ∞), depending on the specific PAPR regions given in Proposition 3.

 figure: Fig. 7

Fig. 7 DCO-OFDM PAPR CCDF versus PAPR with N = 1024.

Download Full Size | PDF

Figure 8 displays the PAPR CCDF of DCO-OFDM versus the clipping bounds l and u, respectively. From Figs. 8(b) and 8(a), one can observe that the CCDF is increasing in l and decreasing in u within some region, which is consistent with Proposition 2. Note that such a monotonicity only holds for the PAPR within [0, θmin) or [θmin, θmax). When l or u changes, the dividing points θmin and θmax also change. Hence, a given PAPR may move out the region [0, θmin) or [θmin, θmax), which leads the discontinuity in Figs. 8(a) and 8(b).

 figure: Fig. 8

Fig. 8 DCO-OFDM PAPR CCDF versus l and u with N = 1024. In (a) u = 5 and in (b) l = −5.

Download Full Size | PDF

In Fig 9, we plot the simulated and analytical PAPR CCDFs of the clipped ACO-OFDM, PAM-DMT, and Filp-OFDM signals with N = 1024 subcarriers. Given the same upper clipping bound u, the clipped ACO-OFDM, PAM-DMT, and Filp-OFDM signals have the same PAPR ceiling θ0,U, which is increasing in u, i.e., the more the signal is clipped above, the smaller the maximum PAPR is. The PAPR CCDFs of ACO-OFDM and PAM-DMT are nearly identical and smaller than that of Flip-OFDM. This is consistent with the relation in (19).

 figure: Fig. 9

Fig. 9 ACO-OFDM, PAM-DMT, Flip-OFDM PAPR CCDFs versus PAPR with N = 1024.

Download Full Size | PDF

In Fig. 10, we compare the PAPR CCDFs of the clipped DCO-OFDM, ACO-OFDM, PAM-DMT, and Flip-OFDM signals. Given the same upper clipping bound u, the PAPR CCDF of the half-clipped (l = 0) DCO-OFDM signal is quite similar to that of the clipped ACO-OFDM or PAM-DMT signal (therefore their PAPR CCDF curves overlap in Fig. 10), while Flip-OFDM has the largest PAPR CCDF. However, the DCO-OFDM signal is very unlikely to be half clipped from below, which will cause a great deal of information loss. Therefore, in practice, we have l < 0 for DCO-OFDM and in this case, from Figs. 10(a)–10(c), DCO-OFDM has a smaller PAPR than ACO-OFDM, PAM-DMT, and Flip-OFDM. From Figs. 10(c) and 10(d), one can observe that when the normalized clipping bound is equal to 5, the effect of upper or lower clipping tends to saturation.

 figure: Fig. 10

Fig. 10 Comparison of the PAPR CCDFs of the four VLC OFDM schemes.

Download Full Size | PDF

5. Conclusion

We investigated PAPR distributions of the clipped DCO-OFDM, ACO-OFDM, PAM-DMT, and Flip-OFDM signals and took into account both lower and upper clippings. The PAPR CCDFs of the four VLC OFDM schemes were analytically characterized and shown to be piecewise functions. We further investigated the monotonic properties of the PAPR CCDF and the PAPR ceiling with respect to the lower and upper clipping bounds, which lead to useful insights for VLC OFDM systems. The derived analytical results were verified by numerical simulations.

Appendix

A1. Proof of theorem 1

From (9) we know E[|sclip,n|2]=σs2c(l,u). The CDF of the PAPR is

P(PAPRx)=P(max0nN1|sclip,n|2σs2c(l,u)x)=P(|sclip,n|2σs2c(l,u)x,n=0,,N1)
Although the Hermitian symmetry introduces correlation among sn, for a large N the correlation tends to be zero [21], i.e., sn (sclip,n) is independent for n = 0, . . . , N − 1. Hence, we have
P(PAPRx)=[P(|sclip,n|2σs2c(l,u)x)]N.
Consider the following four situations: 1) If σsc(l,u)x<min{BL,BU}, then
P(|sclip,n|2σs2c(l,u)x)=P(σsc(l,u)xsclip,nσsc(l,u)x)=σsc(l,u)xσsc(l,u)xfdco(w)dw=2Φ(c(l,u)x)1.
2) If σsc(l,u)xmax{BL,BU}, then
P(|sclip,n|2σs2c(l,u)x)=P(BLsclip,nBU)=BLBUfdco(w)dw=1.
3) If BL=min{BL,BU}σsc(l,u)x<max{BL,BU}=BU, then
P(|sclip,n|2σs2c(l,u)x)=P(BLsclip,nσsc(l,u)x)=BLσsc(l,u)xfdco(w)dw=Φ(c(l,u)x).
4) If BU=min{BL,BU}σsc(l,u)x<max{BL,BU}=BL, then
P(|sclip,n|2σs2c(l,u)x)=P(σsc(l,u)xsclip,nBU)=σsc(l,u)xBUfdco(w)dw=Φ(c(l,u)x).
Therefore, the CDF of the PAPR is given by
P(PAPRx)={[2Φ(c(l,u)x1)]N,x<θmin[Φ(c(l,u)x)]N,θminx<θmax1,xθmax.
The CCDF in (11) is 1 − P(PAPRx).

A2. Proof of proposition 1

It is easily seen that θL is decreasing in u, since c(l, u) is increasing in u. For l, we have

θLl=2lc(l,u)l22lΦ(l)c2(l,u)=2l(c(l,u)l2Φ(l))c2(l,u)=2lψL(l,u)c2(l,u)
where ψL (l, u) = c(l, u) − l2Φ(l), which is increasing u. It follows that
ψL(l,u)l=2lΦ(l)2lΦ(l)l2g(l)=l2g(l)0,
implying that ψL (l, u) is decreasing in l. So we have ψL (l, u) ≥ ψL (0, 0) = c(0, 0) = 0 and ∂θL/∂l ≤ 0. Therefore, θL is decreasing in l.

Similarly, θU is increasing in l, since c(l, u) is decreasing in l. For u, we have

θUu=2uc(l,u)u22u(1Φ(u))c2(l,u)=2u(c(l,u)u2+u2Φ(u))c2(l,u)=2uψU(l,u)c2(l,u)
where ψU (l, u) = c(l, u) − u2 + u2Φ(u), which is decreasing in l. It follows that
ψU(l,u)u=2u(1Φ(u))2u+2uΦ(u)+u2g(u)=u2g(u)0,
implying that ψU (l, u) is increasing in u. So we have ψU (l, u) ≥ ψU (0, 0) = c(0, 0) = 0 and ∂θU/∂u ≥ 0. Therefore, θU is increasing in u.

Now, consider the symmetric case where −l = u and θ = u2/c(−u, u). It can be verified that dc(−u, u)/du = 4u(1 − Φ(u)). We have

dθdu=2uc(u,u)u24u(1Φ(u))c2(u,u)=2u(2Φ(u)2ug(u)1)c2(u,u)=2uψ(u)c2(u,u)
where ψ(u) = 2Φ(u) − 2ug(u) − 1. It follows that ψ′(u) = 2u2g(u) ≥ 0, implying that ψ(u) is increasing in u and ψ(u) ≥ ψ(0) = 0. Therefore, dθ/du ≥ 0 and θ is increasing in u.

A3. Proof of proposition 3

If xθmax, then Gdco(x) = 0 ≤ Gunc(x). If x < θmin, then Gdco(x)=1[2Φ(c(l,u)x)1]N. In this case, since c(l, u) ≤ c(−∞, +∞) = 1, we obtain Φ(c(l,u)x)Φ(x) and thus Gdco(x) ≥ Gunc(x). If θminx < θmax, then Gdco(x)=1[Φ(c(l,u)x)]N. In this case, Gdco(x) ≤ Gunc(x) if Φ(c(l,u)x)2Φ(x)1, otherwise Gdco(x) ≥ Gunc(x).

Consider the symmetric clipping. If xθ, then Gsym(x) = 0 ≤ Gunc(x). If 0 ≤ x < θ, then Gsym(x)=1[2Φ(c(u,u)x)1]N. In this case, since c(−u, u) ≤ 1, we obtain Φ(c(u,u)x)Φ(x) and thus Gsym(x) ≥ Gunc(x).

A4. Proof of theorem 2

From Section 3.2, we know E[|sclip,n|2]=σs2c(0,u). The CDF of the ACO-OFDM PAPR is

P(PAPRx)=P(max0nN1|sclip,n|2σs2c(0,u)x)=P(sclip,nσsc(0,u)x,n=0,,N1)
where we have used the fact that sclip,n is nonnegative. It follows from (4) that sn+N2=sn for n=0,,N21, i.e., sn and sn+N2 are correlated, so are sclip,n and sclip,n+N2. On the other hand, for a large N, sn (sclip,n) is independent for n=0,,N21. Therefore, we have
P(PAPRx)=n=0N21P(sclip,nσsc(0,u)x,sclip,n+N2σsc(0,u)x).
It follows that
P(sclip,nσsc(0,u)x,sclip,n+N2σsc(0,u)x)=P(sclip,nσsc(0,u)x,sclip,n+N2σsc(0,u)x,sn0)+P(sclip,nσsc(0,u)x,sclip,n+N2σsc(0,u)x,sn<0)(a)=P(sclip,nσsc(0,u)x,sn0)+P(sclip,n+N2σsc(0,u)x,sn<0)(b)=2P(sclip,nσsc(0,u)x,sn0)
where (a) is because if sn ≥ 0, then sclip,n+N2=0 and if sn < 0, then sclip,n = 0, and (b) is because sclip,n and sclip,n+N2 have the same distribution and P(sn ≥ 0) = P(sn < 0) = 0.5. If σsc(0,u)xBU, then P(sclip,nσsc(0,u)x,sn0)=P(sn0)=0.5. If σsc(0,u)x<BU, then
P(sclip,nσsc(0,u)x,sn0)=P(0snσsc(0,u)x)=0σsc(0,u)x12πσsew22σs2dw=Φ(c(0,u)x)0.5.
Consequently, the CDF of the PAPR is
P(PAPRx)={[2Φ(c(0,u)x)1]N2,0x<θ0,U1,xθ0,U
which leads to the CCDF in (16).

A5. Proof of theorem 4

It follows from Section 3.2 that E[sflip,n2]=σs2c(0,u). The CDF of the PAPR is

P(PAPRx)=P(sclip,nσsc(0,u)x,n=0,,2N1)=n=0N1P(sclip,nσsc(0,u)x,sclip,n+Nσsc(0,u)x)
where we have used the facts that sclip,n is nonnegative and for a large N, sn(sclip,n) is independent for n = 0, . . . , N − 1, but sclip,n and sclip,n+N are correlated.
P(sclip,nσsc(0,u)x,sclip,n+Nσsc(0,u)x)=P(sclip,nσsc(0,u)x,sclip,n+Nσsc(0,u)x,sn0)+P(sclip,nσsc(0,u)x,sclip,n+Nσsc(0,u)x,sn<0)(a)=P(sclip,nσsc(0,u)x,sn0)+P(sclip,n+Nσsc(0,u)x,sn<0)(b)=2P(sclip,nσsc(0,u)x,sn0)
where (a) is because if sn ≥ 0, then sclip,n+N = 0 and if sn < 0, then sclip,n = 0, and (b) is because sclip,n and sclip,n+N have the same distribution and P(sn ≥ 0) = P(sn < 0) = 0.5. Then, following the similar steps in Appendix A4, we can obtain the CCDF in (18).

Funding

973 Program of China (2013CB329204); National Natural Science Foundation of China (61571107 and 61521061); the Natural Science Foundation of Jiangsu Province (BK20160069); The Alexander von Humboldt Foundation; Fundamental Research Funds for the Central Universities; EPSRC (EP/N004558/1 and EP/N023862/1).

References and links

1. H. Elgala, R. Mesleh, and H. Haas, “Indoor optical wireless communication: Potential and state-of-the-art,” IEEE Commun. Mag. 49(9), 56–62 (2011). [CrossRef]  

2. T. Komine and M. Nakagawa, “Fundamental analysis for visible-light communication system using LED lights,” IEEE Trans. Consumer Electron. 50(1), 100–107 (2004). [CrossRef]  

3. R. Zhang, J. Wang, Z. Wang, Z. Xu, L. Hanzo, and C. Zhao, “Visible light communications in heterogeneous networks: Paving the way for user-centric design,” IEEE Wireless Commun. , 22(2), 8–16 (2015). [CrossRef]  

4. Y. Tao, X. Liang, J. Wang, and C. Zhao, “Scheduling for indoor visible light communication based on graph theory,” Optical Express , 23(3), 2737–2752, (2015). [CrossRef]  

5. L. Chen, J. Wang, J. Zhou, D. W. K. Ng, R. Schober, and C. Zhao, “Distributed user-centric scheduling for visible light communication networks,” Optical Express , 24(14), 15570–15589, (2016). [CrossRef]  

6. J. Armstrong, “OFDM for optical communications,” IEEE J. Lightw. Technol. 27(3), 189–204 (2009). [CrossRef]  

7. J. Armstrong and A. Lowery, “Power efficient optical OFDM,” Electron. Lett. 42(6), 370–372 (2006). [CrossRef]  

8. S. C. J. Lee, S. Randel, F. Breyer, and A. M. J. Koonen, “PAM-DMT for intensity-modulated and direct-detection optical communication systems,” IEEE Photon. Technol. Lett. 21(23), 1749–1751 (2009). [CrossRef]  

9. N. Fernando, Y. Hong, and E. Viterbo, “Flip-OFDM for unipolar communication systems,” IEEE Trans. Commun. 60(12), 3726–3733 (2012). [CrossRef]  

10. S. H. Han and J. H. Lee, “An overview of peak-to-average power ratio reduction techniques for multicarrier transmission,” IEEE Wireless Commun. 12(2), 56–65 (2005). [CrossRef]  

11. H. Ochiai and H. Imai, “On the distribution of the peak-to-average power ratio in OFDM signals,” IEEE Trans. Commun. 49(2), 282–289 (2001). [CrossRef]  

12. N. Jacklin and Z. Ding, “A linear programming based tone injection algorithm for PAPR reduction of OFDM and linearly precoded systems,” IEEE Trans. Circuits and Syst. I: Reg. Papers 60(7), 1937–1945 (2013). [CrossRef]  

13. W. O. Popoola, Z. Ghassemlooy, and B. G. Stewart, “Pilot-assisted PAPR reduction technique for optical OFDM communication systems,” IEEE J. Lightw. Technol. 32(7), 1374–1382 (2014). [CrossRef]  

14. K. Ying, Z. Yu, R. J. Baxley, and G. T. Zhou, “Constrained clipping for PAPR reduction in VLC systems with dimming control,” in Proc. IEEE GlobalSIP (IEEE, 2015), pp. 1327–1331.

15. C. Ma, H. Zhang, M. Yao, Z. Xu, and K. Cui, “Distributions of PAPR and crest factor of OFDM signals for VLC,” in Proc. IEEE Photonics Society Summer Topical Meeting Series (IEEE, 2012), pp. 119–120.

16. Z. Yu, R. J. Baxleyt, and G. T. Zhou, “Distributions of upper PAPR and lower PAPR of OFDM signals in visible light communications,” in Proc. IEEE ICASSP (IEEE, 2014), pp. 355–359.

17. S. Dimitrov, S. Sinanovic, and H. Haas, “Clipping noise in OFDM-based optical wireless communication systems,” IEEE Trans. Commun. 60(4), 1072–1081 (2012). [CrossRef]  

18. K. Ying, Z. Yu, R. J. Baxley, H. Qian, G.-K. Chang, and G. T. Zhou, “Nonlinear distortion mitigation in visible light communications,” IEEE Wireless Commun. , 22(2), 36–45 (2015). [CrossRef]  

19. K. Ying, Z. Yu, R. J. Baxley, and G. T. Zhou, “Optimization of signal-to-noise-plus-distortion ratio for dynamic-rangelimited nonlinearities,” Digital Signal Process. , 36, 104–114 (2015). [CrossRef]  

20. X. Ling, J. Wang, X. Liang, Z. Ding, and C. Zhao, “Offset and power optimization for DCO-OFDM in visible light communication systems,” IEEE Trans. Signal Process. , 64(2), 349–363 (2016). [CrossRef]  

21. H. Yu, M. Chen, and G. Wei, “Distribution of PAR in DMT systems,” Electron. Lett. 39(10), 799–801 (2003). [CrossRef]  

22. N. Johnson, S. Kotz, and N. Balakrishnan, Continuous Univariate Distributions (John Wiley, 1994) 2nd ed.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 Transmitter block diagram of the DCO-OFDM VLC System.
Fig. 2
Fig. 2 Transmitter block diagram of the ACO-OFDM VLC System.
Fig. 3
Fig. 3 Transmitter block diagram of the PAM-DMT VLC System.
Fig. 4
Fig. 4 Transmitter block diagram of the Flip-OFDM VLC System.
Fig. 5
Fig. 5 (a) values of c(l, u) and (b) values of θL and θU for different l and u.
Fig. 6
Fig. 6 PAPR CCDFs of the four VLC OFDM schemes with different subcarrier numbers.
Fig. 7
Fig. 7 DCO-OFDM PAPR CCDF versus PAPR with N = 1024.
Fig. 8
Fig. 8 DCO-OFDM PAPR CCDF versus l and u with N = 1024. In (a) u = 5 and in (b) l = −5.
Fig. 9
Fig. 9 ACO-OFDM, PAM-DMT, Flip-OFDM PAPR CCDFs versus PAPR with N = 1024.
Fig. 10
Fig. 10 Comparison of the PAPR CCDFs of the four VLC OFDM schemes.

Equations (50)

Equations on this page are rendered with MathJax. Learn more.

S k = S N k * , k = 1 , , N 2 1
S = [ 0 , S 1 , S 2 , , S N 2 1 , 0 , S N 2 1 * , , S 2 * , S 1 * ] .
s n = 1 N k = 0 N 1 S k e j 2 π k n N = 2 N k = 1 N 2 1 Re { S k e j 2 π k n N }
s clip , n = Clip [ s n ] = { B U , s n B U s n , B L < s n < B U B L , s n B L
S = [ 0 , S 1 , 0 , , S N 2 1 , 0 , S N 2 1 * , , 0 , S 1 * ] .
s n = 1 N k = 0 N 1 S k e j 2 π k n N = 2 N k = 1 N 2 1 Re { S k e j 2 π k n N } = 2 N t = 1 N / 4 Re { S 2 t 1 e j 2 π ( 2 t 1 ) n N } .
s n + N 2 = 2 N t = 1 N / 4 Re { S 2 t 1 e j 2 π ( 2 t 1 ) ( n + N 2 ) N } = s n , n = 0 , , N 2 1 .
s clip , n = Clip [ s n ] = { B U , s n B U s n , 0 < s n < B U 0 , s n 0 .
S = [ 0 , j A 1 , j A 2 , , j A N 2 1 , 0 , j A N 2 1 , , j A 2 , j A 1 ] .
s n = 1 N k = 0 N 1 S k e j 2 π k n N = 1 N k = 1 N 2 1 [ j A k e j 2 π k n N j A k e j 2 π ( N k ) n N ] = 2 N k = 1 N 2 1 A k sin 2 π k n N
s N n = 2 N k = 1 N 2 1 A k sin 2 π k ( N n ) N = s n n = 1 , , N 2 1
s n + = { s n , if s n 0 0 , otherwise and s n = { s n , if s n < 0 0 , otherwise .
s clip , n = Clip [ s ¯ n ] = { B U , s n B U s n , s n < B U
PAPR = max 0 n N 1 | s clip , n | 2 E [ | s clip , n | 2 ]
l = B L σ s and u = B U σ s .
f dco ( w ) = Φ ( l ) δ ( w + B L ) + 1 2 π σ s e w 2 2 σ s 2 [ u ( w + B L ) u ( w B U ) ] + [ 1 Φ ( u ) ] δ ( w B U ) .
f aco ( w ) = 1 2 δ ( w ) + 1 2 π σ s e w 2 2 σ s 2 [ u ( w ) u ( w B U ) ] + [ 1 Φ ( u ) ] δ ( w B U ) .
P dco ele = E [ | s clip , n | 2 ] = E [ s clip , n 2 ] = + w 2 f dco ( w ) d w = σ s 2 c ( l , u )
c ( l , u ) = ( l 2 1 ) Φ ( l ) ( u 2 1 ) Φ ( u ) + u 2 + lg ( l ) u g ( u ) .
P dco opt = E [ s clip , n ] + B L = + w f dco ( w ) d w = σ s [ l Φ ( l ) + g ( l ) g ( u ) + u ( 1 Φ ( u ) ) ] + B L
F dco ( x ) = P ( PAPR > x ) = { 1 [ 2 Φ ( c ( l , u ) x ) 1 ] N , 0 x < θ min 1 [ Φ ( c ( l , u ) x ) ] N , θ min x < θ max 0 , x θ max
θ L = l 2 c ( l , u ) and θ U = u 2 c ( l , u ) .
F sym ( x ) = { 1 , [ 2 Φ ( c ( u , u ) x ) 1 ] N , 0 x < θ 0 , x θ
F unc ( x ) = 1 [ 2 Φ ( x ) 1 ] N , x 0 .
F dco ( x ) { F unc ( x ) , x < θ min or β x < θ max F unc ( x ) , x θ max or θ min x α
F sym ( x ) { F unc ( x ) , 0 x < θ F unc ( x ) , x θ .
F aco ( x ) = P ( PAPR > x ) = { 1 [ 2 Φ ( c ( 0 , u ) x ) 1 ] N 2 , 0 x < θ 0 , U 0 , x θ 0 , U
F pam ( x ) = P ( PAPR > x ) = { 1 [ 2 Φ ( c ( 0 , u ) x 1 ] N 2 1 , 0 x < θ 0 , U 0 , x θ 0 , U
PAPR = max 0 n 2 N 1 | s clip , n | 2 E [ | s clip , n | 2 ]
F flip ( x ) = P ( PAPR > x ) = { 1 [ 2 Φ ( c ( 0 , u ) x 1 ) ] N , 0 x < θ 0 , U 0 , x θ 0 , U
F aco ( x ) F pam ( x ) F flip ( x )
P ( PAPR x ) = P ( max 0 n N 1 | s clip , n | 2 σ s 2 c ( l , u ) x ) = P ( | s clip , n | 2 σ s 2 c ( l , u ) x , n = 0 , , N 1 )
P ( PAPR x ) = [ P ( | s clip , n | 2 σ s 2 c ( l , u ) x ) ] N .
P ( | s clip , n | 2 σ s 2 c ( l , u ) x ) = P ( σ s c ( l , u ) x s clip , n σ s c ( l , u ) x ) = σ s c ( l , u ) x σ s c ( l , u ) x f dco ( w ) d w = 2 Φ ( c ( l , u ) x ) 1 .
P ( | s clip , n | 2 σ s 2 c ( l , u ) x ) = P ( B L s clip , n B U ) = B L B U f dco ( w ) d w = 1 .
P ( | s clip , n | 2 σ s 2 c ( l , u ) x ) = P ( B L s clip , n σ s c ( l , u ) x ) = B L σ s c ( l , u ) x f dco ( w ) d w = Φ ( c ( l , u ) x ) .
P ( | s clip , n | 2 σ s 2 c ( l , u ) x ) = P ( σ s c ( l , u ) x s clip , n B U ) = σ s c ( l , u ) x B U f dco ( w ) d w = Φ ( c ( l , u ) x ) .
P ( PAPR x ) = { [ 2 Φ ( c ( l , u ) x 1 ) ] N , x < θ min [ Φ ( c ( l , u ) x ) ] N , θ min x < θ max 1 , x θ max .
θ L l = 2 l c ( l , u ) l 2 2 l Φ ( l ) c 2 ( l , u ) = 2 l ( c ( l , u ) l 2 Φ ( l ) ) c 2 ( l , u ) = 2 l ψ L ( l , u ) c 2 ( l , u )
ψ L ( l , u ) l = 2 l Φ ( l ) 2 l Φ ( l ) l 2 g ( l ) = l 2 g ( l ) 0 ,
θ U u = 2 u c ( l , u ) u 2 2 u ( 1 Φ ( u ) ) c 2 ( l , u ) = 2 u ( c ( l , u ) u 2 + u 2 Φ ( u ) ) c 2 ( l , u ) = 2 u ψ U ( l , u ) c 2 ( l , u )
ψ U ( l , u ) u = 2 u ( 1 Φ ( u ) ) 2 u + 2 u Φ ( u ) + u 2 g ( u ) = u 2 g ( u ) 0 ,
d θ d u = 2 u c ( u , u ) u 2 4 u ( 1 Φ ( u ) ) c 2 ( u , u ) = 2 u ( 2 Φ ( u ) 2 u g ( u ) 1 ) c 2 ( u , u ) = 2 u ψ ( u ) c 2 ( u , u )
P ( PAPR x ) = P ( max 0 n N 1 | s clip , n | 2 σ s 2 c ( 0 , u ) x ) = P ( s clip , n σ s c ( 0 , u ) x , n = 0 , , N 1 )
P ( PAPR x ) = n = 0 N 2 1 P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N 2 σ s c ( 0 , u ) x ) .
P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N 2 σ s c ( 0 , u ) x ) = P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N 2 σ s c ( 0 , u ) x , s n 0 ) + P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N 2 σ s c ( 0 , u ) x , s n < 0 ) ( a ) = P ( s clip , n σ s c ( 0 , u ) x , s n 0 ) + P ( s clip , n + N 2 σ s c ( 0 , u ) x , s n < 0 ) ( b ) = 2 P ( s clip , n σ s c ( 0 , u ) x , s n 0 )
P ( s clip , n σ s c ( 0 , u ) x , s n 0 ) = P ( 0 s n σ s c ( 0 , u ) x ) = 0 σ s c ( 0 , u ) x 1 2 π σ s e w 2 2 σ s 2 d w = Φ ( c ( 0 , u ) x ) 0.5 .
P ( PAPR x ) = { [ 2 Φ ( c ( 0 , u ) x ) 1 ] N 2 , 0 x < θ 0 , U 1 , x θ 0 , U
P ( PAPR x ) = P ( s clip , n σ s c ( 0 , u ) x , n = 0 , , 2 N 1 ) = n = 0 N 1 P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N σ s c ( 0 , u ) x )
P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N σ s c ( 0 , u ) x ) = P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N σ s c ( 0 , u ) x , s n 0 ) + P ( s clip , n σ s c ( 0 , u ) x , s clip , n + N σ s c ( 0 , u ) x , s n < 0 ) ( a ) = P ( s clip , n σ s c ( 0 , u ) x , s n 0 ) + P ( s clip , n + N σ s c ( 0 , u ) x , s n < 0 ) ( b ) = 2 P ( s clip , n σ s c ( 0 , u ) x , s n 0 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.