fair resource allocation schemes for cooperative dynamic

13
Fair Resource Allocation Schemes for Cooperative Dynamic Free-Space Optical Networks Abdallah S. Ghazy, Hossam A. I. Selmy, and Hossam M. H. Shalaby AbstractFree-space optics (FSO) presents a promising solution for the last-mile connectivity problem. However, adverse weather conditions like fog, rain, and snow, can significantly degrade the performance of FSO links. Thus, the implementation of reliable FSO networks becomes a significant issue. At the same implementation cost, dynamic (reconfigurable) FSO networks are more robust against severe weather conditions over static ones. In this paper, two resource allocation schemes are proposed for dynamic cooperative FSO networks. Each scheme is formulated as a multi-objective optimization problem, where capacity, reli- ability-fairness, average transmitted power, and bit-error rate functions are targeted. One scheme prioritizes the reliability-fairness over capacity, while the other does the opposite. The simulation results reveal that the proposed schemes are more reliable and cost efficient than tradi- tional relayed networks, especially under severe weather conditions. Index TermsDynamic networks; Free space optics; Lexicographic; Lex-max-min theory; Relayed networks; Resource allocation; Weighted sum. I. INTRODUCTION F ree-space optics (FSO) is a line-of-site (LOS) wireless optical communication technique that is used as a promising and feasible solution for the last-mile connectiv- ity problem, where remote network nodes are connected to a central backbone node. With the significant development in optical technology in the past decade, more FSO links are deployed in a given service area to meet increased de- mands on Internet services and applications [1]. Generally, FSO replaces optical fibers when short implementation time, flexible installation, and low implementation cost are required [2]. As indicated in Fig. 1, an FSO link is used to connect different remote nodes, such as mobile base sta- tions, telephone offices, or private networks, to a central backbone node. In spite of the attractive features of FSO, it suffers from free-space channel impairments in the infrared (IR) band spectrum, which are weather conditions, background radi- ation, and atmospheric turbulence [1,3,4]. Weather condi- tions include fog, rain, and snow that can absorb and scatter the transmitted optical signal [5]. On the other hand, eye safety regulation restricts the amount of average transmitted optical power to a certain threshold, which, in turn, limits the performance and distance of FSO links. To overcome these limitations, suitable network topologies are being investigated to provide the required quality of service (QoS) for different FSO nodes. A conventional FSO network implements static direct links (D-Ls) between a fiber backbone node and FSO nodes, as indicated in Fig. 2(A). Although this static topology has simple and low-cost implementation, it has the worst communication performance against severe weather condi- tions. To overcome this degradation, serial-relayed topol- ogy is addressed [5]. In this topology, one or more relays are inserted between far nodes and the backbone node. The relay has two optical transceivers and is located at equal distances from other nodes (optimal placement) [6], as indicated in Fig. 2(B) for partial relayed link net- works (P-Ls). By increasing the number of intermediate re- lays between remote nodes and the backbone node, the best FSO link performance can be achieved. Obviously, this en- hancement in network performance comes at a significant increase in network cost. The topology, where each node is supported by one relay, is called fully relayed links (F-Ls), as indicated in Fig. 2(C). A more robust static FSO network is achieved by imple- menting mesh topology [1,7]. In this topology, redundant FSO links are used between users to keep connectivities to the backbone node as indicated in Fig. 2(D). Clearly, at given atmospheric conditions, each user selects the path that achieves the highest transmission rate at an accept- able bit-error rate. Obviously, enhancing performance of static FSO networks requires installation of a large num- ber of redundant optical transceivers, which, in turn, raises the network cost. Better performance can be achieved at much lower cost by implementing dynamic (reconfigurable) FSO network topologies [8]. These dynamic topologies are classified according to resource sharing into cooperative and non- cooperative topologies. In dynamic non-cooperative FSO http://dx.doi.org/10.1364/JOCN.8.000822 Manuscript received December 7, 2015; revised June 29, 2016; accepted August 31, 2016; published October 7, 2016 (Doc. ID 255159). A. S. Ghazy is with the Department of Electronics and Communications Engineering, Egypt-Japan University of Science and Technology (E-JUST), Alexandria 21934, Egypt (e-mail: [email protected]). H. A. I. Selmy is with the National Institute of Laser Enhanced Science (NILES), Cairo University, Egypt. H. M. H. Shalaby is with the Electrical Engineering Department, Alexandria University, Alexandria 21544, Egypt. 822 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al. 1943-0620/16/110822-13 Journal © 2016 Optical Society of America

Upload: others

Post on 10-Dec-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fair Resource Allocation Schemes for Cooperative Dynamic

Fair Resource Allocation Schemesfor Cooperative Dynamic Free-Space

Optical NetworksAbdallah S. Ghazy, Hossam A. I. Selmy, and Hossam M. H. Shalaby

Abstract—Free-space optics (FSO) presents a promisingsolution for the last-mile connectivity problem. However,adverse weather conditions like fog, rain, and snow, cansignificantly degrade the performance of FSO links. Thus,the implementation of reliable FSO networks becomes asignificant issue. At the same implementation cost, dynamic(reconfigurable) FSO networks are more robust againstsevere weather conditions over static ones. In this paper,two resource allocation schemes are proposed for dynamiccooperative FSO networks. Each scheme is formulated asamulti-objective optimization problem, where capacity, reli-ability-fairness, average transmitted power, and bit-errorrate functions are targeted. One scheme prioritizes thereliability-fairness over capacity, while the other does theopposite. The simulation results reveal that the proposedschemes are more reliable and cost efficient than tradi-tional relayed networks, especially under severe weatherconditions.

Index Terms—Dynamic networks; Free space optics;Lexicographic; Lex-max-min theory; Relayed networks;Resource allocation; Weighted sum.

I. INTRODUCTION

F ree-space optics (FSO) is a line-of-site (LOS) wirelessoptical communication technique that is used as a

promising and feasible solution for the last-mile connectiv-ity problem, where remote network nodes are connected toa central backbone node. With the significant developmentin optical technology in the past decade, more FSO linksare deployed in a given service area to meet increased de-mands on Internet services and applications [1]. Generally,FSO replaces optical fibers when short implementationtime, flexible installation, and low implementation costare required [2]. As indicated in Fig. 1, an FSO link is usedto connect different remote nodes, such as mobile base sta-tions, telephone offices, or private networks, to a centralbackbone node.

In spite of the attractive features of FSO, it suffers fromfree-space channel impairments in the infrared (IR) bandspectrum, which are weather conditions, background radi-ation, and atmospheric turbulence [1,3,4]. Weather condi-tions include fog, rain, and snow that can absorb andscatter the transmitted optical signal [5]. On the otherhand, eye safety regulation restricts the amount of averagetransmitted optical power to a certain threshold, which, inturn, limits the performance and distance of FSO links.To overcome these limitations, suitable network topologiesare being investigated to provide the required quality ofservice (QoS) for different FSO nodes.

A conventional FSO network implements static directlinks (D-Ls) between a fiber backbone node and FSO nodes,as indicated in Fig. 2(A). Although this static topologyhas simple and low-cost implementation, it has the worstcommunication performance against severe weather condi-tions. To overcome this degradation, serial-relayed topol-ogy is addressed [5]. In this topology, one or more relaysare inserted between far nodes and the backbone node.The relay has two optical transceivers and is located atequal distances from other nodes (optimal placement)[6], as indicated in Fig. 2(B) for partial relayed link net-works (P-Ls). By increasing the number of intermediate re-lays between remote nodes and the backbone node, the bestFSO link performance can be achieved. Obviously, this en-hancement in network performance comes at a significantincrease in network cost. The topology, where each node issupported by one relay, is called fully relayed links (F-Ls),as indicated in Fig. 2(C).

A more robust static FSO network is achieved by imple-menting mesh topology [1,7]. In this topology, redundantFSO links are used between users to keep connectivitiesto the backbone node as indicated in Fig. 2(D). Clearly,at given atmospheric conditions, each user selects the paththat achieves the highest transmission rate at an accept-able bit-error rate. Obviously, enhancing performance ofstatic FSO networks requires installation of a large num-ber of redundant optical transceivers, which, in turn, raisesthe network cost.

Better performance can be achieved at much lower costby implementing dynamic (reconfigurable) FSO networktopologies [8]. These dynamic topologies are classifiedaccording to resource sharing into cooperative and non-cooperative topologies. In dynamic non-cooperative FSOhttp://dx.doi.org/10.1364/JOCN.8.000822

Manuscript received December 7, 2015; revised June 29, 2016; acceptedAugust 31, 2016; published October 7, 2016 (Doc. ID 255159).

A. S. Ghazy is with the Department of Electronics and CommunicationsEngineering, Egypt-Japan University of Science and Technology (E-JUST),Alexandria 21934, Egypt (e-mail: [email protected]).

H. A. I. Selmy is with the National Institute of Laser Enhanced Science(NILES), Cairo University, Egypt.

H. M. H. Shalaby is with the Electrical Engineering Department,Alexandria University, Alexandria 21544, Egypt.

822 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

1943-0620/16/110822-13 Journal © 2016 Optical Society of America

Page 2: Fair Resource Allocation Schemes for Cooperative Dynamic

network topologies, no resources are shared among differ-ent users [8]. Users with bad links switch their traffic tousers with relatively better links and the transmission rateof each user is kept the same. This is achieved by increasingtransmission rates of good links by the sum of switchedrates. However, increasing the transmission rate of an op-tical link is not feasible and has practical limitations [3].

To cope with these limitations, dynamic cooperativetopologies have been introduced. In these topologies, nodeswith bad optical links switch their transmissions to nodeswith better links, and the transmission rate of a good link isdivided between node traffic and switched transmissions tomaintain QoS for the switched ones. In other words, thenetwork nodes cooperate and share their resources (opti-cal bandwidths) to maintain connectivities between back-bone node and far nodes under severe weather conditions.

Moreover, at given weather conditions, the network resour-ces can be fairly allocated (achieve nearly the same trans-mission rate as the backbone node) among different usersby implementing a proper resource allocation scheme.Although the number of optical transceivers available ateach node plays an important role in the network perfor-mance, it is still much lower than the number requiredin static topologies to achieve the same performance.

Two resource allocation schemes have recently been intro-duced to enhance performance of dynamic cooperative FSOnetworks against various weather conditions [9–11]. In thispaper, the impacts of power level adaptation, operating wave-lengths, and modulation techniques on the performance ofthese resource allocation schemes are investigated. Also,a comparison with other resource allocation schemes thatuse weighted sum optimization is carried out. The rest of thispaper is organized as follows. Section II presents theFSO linkmodel. Section III illustrates reconfigurable cooperative FSOnetwork parameters. Section IV introduces the proposedresource allocation schemes. Section V shows the numericalevaluations for proposed resource allocation schemes. Last,Section VI is the conclusion and remarkable notes.

II. FSO LINK MODEL

Two main factors affect FSO link performance, namely,link losses and noise. Link losses include both atmosphericand geometric losses. These losses cause signal scattering,absorbing, and spreading. System noise includes both ex-ternal noise (ambient or background noise) and internalnoise (dark current and thermal noise). Also, the selectedmodulation format and wavelength play important roles inFSO link performance [12].

A. Link Losses

FSO channel impairments have both atmospheric andgeometric losses. Atmospheric loss includes fog, rain, snow,and scintillation [4]. In FSO link modeling, the collectingaperture diameter of an optical receiver may be larger thanthe spatial scale of the optical scintillation that is caused byatmospheric turbulence. In this condition, the FSO receiverwill average fluctuations of the received waveform over theaperture area, leading to reduced signal fluctuations, espe-cially if the network experiences weak turbulence [13–15].In this paper, the receiver diameter is assumed to be largeso that the impact of weak turbulence is small [5,16].Under severe weather conditions, the scintillation losshas relatively small attenuation and can be neglected[5,16]. In this case, the total FSO link loss is given by

γ � γfog � γrain � γsnoww� γsnowd

� γgeo; (1)

where γ presents the total link loss in dB. Also, γfog, γrain,γsnoww

, γsnowd, and γgeo are fog, rain, wet snow, dry snow, and

geometric losses, respectively, in dB.

1) Atmospheric Losses: The transmission properties ofoptical signals in free space vary with atmospheric state.

FSO link

Backbone

FSO node

Fig. 1. Last-mile FSO connection; the end users could have wireor wireless connections to an FSO node.

(A) (B)

(C) (D)

Node with more than two transceivers

FSO Link

Relay Node

Node with one transceiver

Fiber Backbone

Fig. 2. FSO static topologies: (A) direct link (D-L), (B) partial re-layed link (P-L), (C) full relayed link (F-L), and (D) partialmesh link.

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 823

Page 3: Fair Resource Allocation Schemes for Cooperative Dynamic

Clearly, particles of space could absorb and scatter thetransmitted optical signal. Light absorption reduces the in-tensity of the light beam and causes attenuation of thereceived optical power. Also, light scattering deflects theincident light from its initial direction, causing spatialspreading [3,4].

Fog loss is defined by several empirical models. For low-visibility (V < 6 km) FSO links, the Kim model is the mostaccurate, and is given by [1]

γfog � 10 × log�exp�β × L��dB; (2)

where β is total extinction coefficient, defined by

β � 3.91V

× �λn550�−Ψ: (3)

Here Ψ is the distribution of scattering particle size and isdescribed by

Ψ �

8>>>><>>>>:

1.6; V > 50 km;1.3; 6 km < V < 50 km;0.16 × V � 0.34; 1 km < V < 6 km;V − 0.5; 0.5 km < V < 1 km;0; V < 0.5 km;

(4)

where V is the visibility in kilometers, λ is the wavelengthin nanometers, and L is the distance in kilometers. In rainyweather, attenuation is caused by optical scattering due todroplets of water. The rain loss is calculated using Japan’sempirical model [4]:

γrain � 1.58 ×D0.63 × L dB; (5)

where D is the rain fall rate in millimeters/hour (mm/h). Insnowy weather, the attenuation is wavelength sensitive,but the sensitivity is not significant, so the loss is approx-imately equal at a wide range of wavelengths, especiallywith dry snow droplets.

The models below are used to calculate wet and dry snowlosses [4]:

a) Wet snow case:

γsnoww� �1.02 × 10−4λ� 3.79� × S0.72

w × L dB; (6)

where Sw is the wet snow fall rate in mm/h.

b) Dry snow case:

γsnowd� �5.42 × 10−5λ� 5.50� × S1.38

d × L dB; (7)

where Sd is the dry snow fall rate in mm/h. Naturally, theseweather phenomena (fog, rain, and snow) rarely occur con-currently and their influences can be studied separately [5].

2) Geometric Loss: Even in clear weather conditions,geometric loss is presented due to optical beam spreadingthrough propagation in free space. This loss is calculatedby [3]

γgeo � 10 × log�dt � L × Θ

dr

�2dB; (8)

where dr is the receiver diameter, dt is the transmitterdiameter (both in millimeters), and Θ is the divergence an-gle in mm.rad/km.

To overcome these losses and maintain network perfor-mance, an FSO link can operate with variable transmittedpower and bit rate levels [17]. Obviously, in clear weatherconditions, the highest transmission rate with an accept-able error rate is achieved by transmitting the lowestaverage power level. In contrast, during bad weather con-ditions, average bit-error rate is maintained by increasingthe level of transmitted power and or reducing the linktransmission rate. However, the maximum average trans-mitted power is restricted by eye safety regulations [1,3,4].

B. FSO System Noise

Two kinds of noise sources are present in FSO links:thermal and shot noise. Thermal noise is generated fromelectronic components of the detection system, while shotnoise is generated from both internal (dark current of theoptical detector) and external sources (random arrival ofsignal and background photons). In some cases, back-ground radiation can even cause link outage because ofthe saturation of the receiver, especially when the fieldof view (FOV) of the receiver is relatively large and/orfaces a radiation source [3,4]. When the background radi-ation level is relatively high, the receiver thermal noisecan be ignored, and the system noise is modeled usinga Poisson model (shot-noise-limited receiver). Hence, atgiven transmitted qt photons/slot and channel loss γ, theprobability of arrival of q photons per slot is given by [12]

Pos�q; qs � qb� ��qs � qb�q

q!× exp�−�qs � qb��; (9)

where qs � γ × qt is the average number of received signalphotons per slot and qb is the average number of receivedambient photons per slot.

C. Modulation Formats

Two prime intensity modulation/direct detection tech-niques, namely, non-return-zero on–off keying (NR-OOK)and pulse position modulation (PPM), are considered inthis paper. PPM enhances receiver sensitivity more thanOOK does, especially at high ambient noise. However, ithas lower bandwidth utilization compared to OOK. Thebit-error rate of OOK, Pe, for a shot-noise-limited receiver,is given by [12]

Pe �12×Xmt

q�0

�qb � qs�q ×exp�−�qb � qs��

q!

� 12×X∞q�mt

�qb�q ×exp�−qb�

q!; (10)

824 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

Page 4: Fair Resource Allocation Schemes for Cooperative Dynamic

where mt is the threshold of bit detection:

mt �qs

log�1� qs

qb

� : (11)

Also, the symbol-error rate of PPM, Ps, for a shot-noise-limited receiver, is given by [12]

Ps � 1 −

1M

× exp��−qs − qb ×M��

X∞q�1

Pos�q; qs � qb� × Xq−1q1�1

Pos�q1; qb�!M−1

XM−1

r�1

�M − 1�!r!�M − r − 1�!�r − 1�

×X∞q�1

Pos�q; qb � qs� × �Pos�q; qb��r

×

0@Xq−1

j�1

Pos�j; qb�1AM−r−1

; (12)

where M is the number of slots per PPM frame. Under aconstraint on the average of transmitted power, PPM out-performs multiple pulse position modulation (MPPM). Incontrast, when a constraint is imposed on the peak of trans-mitted power, MPPM outperforms PPM [16].

The operating wavelength is an important parameterin the design of an FSO link. Indeed, it determines theimpacts of geometric loss, weather attenuations, back-ground radiation, and eye safety on FSO link performance.Generally, longer wavelengths are more immune to chan-nel impairments than shorter ones [1,3,4]. However, in thispaper, homogeneous weather is assumed over all networkregions, i.e., all FSO links are affected by the same atmos-pheric losses and background radiation levels.

III. RECONFIGURABLE COOPERATIVE FSONETWORK PARAMETERS

Currently, the significant innovation in pointing, acquis-ition, and tracking (PAT) systemsmakes dynamic FSO net-works more feasible than before [18]. In reconfigurabletopologies, the number of FSO transceivers can be signifi-cantly reduced by replacing actual FSO relay nodes bytransceivers on other working nodes (virtual FSO relaynodes) [8].

Generally, the cooperative FSO network consists of Nnodes �v1;…; vN� with arbitrary geographical distributionin addition to the backbone node v0. The number of opticaltransceivers at the kth node is denoted by Zk, wherefkg ∈ f1;…; Ng. The backbone node is assumed to beequipped with N optical transceivers. In the FSO networkconsidered, the inner n2 nodes near the backbone node areassumed to have two transceivers, while the far n1 � N −

n2 nodes are assumed to have only one transceiver, i.e.,Zk ∈ f1; 2g. An example of a reconfigurable cooperativeFSO network with one central node and nine remote nodes

is shown in Fig. 3(A). In this network, each node of theinner four nodes has two transceivers (n2 � 4), while eachof the outer five nodes has one transceiver (n1 � 5). In clearweather conditions, all nine nodes are directly connected tothe central node as indicated in Fig. 3(B.1). At severeweather conditions, far nodes could switch their traffic toa neighbor node, as indicated in Fig. 3(B.2).

The losses of all FSO links are summarized in the lossmatrix, γ � �γ00;…; γ0N ;…; γij;…; γN0;…; γNN�, where γij isthe loss coefficient of the FSO link between the transmitterof the ith node and the receiver of the jth node. Clearly,0 ≤ γij ≤ 1, γii � 0, and γij � γji for any fi; jg ∈ f0; 1;…; Ng.At a given weather state, the cooperative FSO networkcan be connected with different feasible configurations thatsatisfy the required QoS parameters, i.e., guarantee mini-mum bit rates at bit-error rates of less than a certainthreshold. The number of these feasible configurations isΛ. For the lth configuration, l ∈ f1;2;…;Λg, the connectionstatus between network nodes is summarized in connec-tion matrix Gl � �gl00;…; gl0N ;…; glij;…; glN0;…; glNN�,where glij is the connection status between the ith andjth nodes in configuration l and glij ∈ f0;1g. The connectionbetween nodes i and j is established in configuration lif glij � 1. Also, bidirectional links are assumed so thatglij � glji and glii � 0.

Clearly, the number of feasible network configurations de-pends on the number of transceivers on each FSO node. Anupper bound for the number of feasible configurations Λ canbe obtained from the symmetric connection matrix Gl withzero diagonal elements. The first row of the upper triangle ofthis matrix represents all connections to/from the backbonenode with different Σi�N

i�0 �Ni � combinations. The next n2 rowsof the triangle represent the connections to/from all nodesthat have two transceivers, resulting in

Qk�n2k�1 Σi�2

i�0�N−ki �

different connections. The next n1 rows of the triangle re-present the connections to/from all nodes that have one

transceiver, resulting inQk�n1

k�1 Σi�1i�0

�N − n2 − ki

�different

connections. Therefore, the size of Λ can be deduced bythe following simple upper bound:

(A)

Node with one transceiver Node with two transceivers

(B.1)

(B.2)

(0)(1)

(6)

(9)

(0)

(1)

(6)

(9)

Fig. 3. Example of a reconfigurable cooperative FSO network.(A) Network topology, and network at various weather conditions:(B.1) clear weather and (B.2) foggy weather.

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 825

Page 5: Fair Resource Allocation Schemes for Cooperative Dynamic

Λ<Xi�N

i�0

�Ni

�×Yk�n2

k�1

Xi�2

i�0

�N−ki

�×Yk�n1

k�1

Xi�1

i�0

�N−n2−ki

�: (13)

Moreover, all FSO links are assumed to have adaptiveaverage transmitted power, i.e., the power of the opticallink between nodes i and j in configuration l is one of dis-crete values, Plij ∈ f0; y1; y2;…; yYg, and y1 < y2 < � � � < yY .The average transmitted power of node k in configurationl is denoted by Plk, where Plk � ΣN

j�0Plkj. However, toincrease link capacity and guarantee an error rate of lessthan a specified maximum BERlij < BERmax, the linkbetween nodes i and j in configuration l adapts its trans-mission rate, Tlij, to be one of X � 1 discrete values, whereTlij ∈ f0; x1; x2…; xXg and x1 < x2 < � � � < xX . The transmis-sion rate of node k in configuration l is denoted by Tlk,where Tlk � ΣN

j�0Tlkj. The bit rate of node k (its own traffic)through connection to node j in configuration l is denotedby Rlkj. The overall bit rate of node k in configuration l isRlk � ΣN

j�0Rlkj. Obviously,Rlk ≤ Tlk, and for practical imple-mentation both Rlk and Tlk ∈ f0; x1; x2;…; xXg. The end-to-end bit-error rate of node k in configuration l, BERlk, isbounded by BERlk ≤ BERmax. The bit rates and bit-errorrates associated with all nodes in the feasible configura-tions can be summarized in (Λ ×N) matrices R and E,respectively. For a given configuration l, the bit rates forall nodes are represented in vector rl (1 ×N), rl ∈ R.Also, the bit-error rates in that configuration are summa-rized in vector el (1 ×N), el ∈ E.

The network capacity and average transmitted power as-sociated with configuration l are given by Cl � ΣN

k�1Rlk andPl � ΣN

k�1Plk, respectively. All capacity and power values as-sociated with all feasible configurations are summarized invectors C �Λ × 1� and P �Λ × 1�, where Cl ∈ C and Pl ∈ P.Also, the maximum network capacity that can be achievedby any configuration is that obtained from direct link con-figuration l and is defined by Cmax � ΣN

k�1Tlk0. Table I

summarizes the assigned symbols for the FSO networkparameters.

IV. PROPOSED FAIR COOPERATIVE RESOURCE

ALLOCATION SCHEMES

Dynamic cooperative FSO networks deploy resource allo-cation schemes to increase capacity, reliability, and fairnessunder various weather conditions. Increasing networkcapacity is achieved by maintaining the largest number ofdirect links to the central node. In addition, increasing net-work reliability implies decreasing the number of droppednodes, while enhancing fairness means nearly the samebit rates are assigned to different nodes. Obviously, underclear weather conditions, all nodes are directly connectedto the central node to get the highest bit rates (maximumnetwork capacity) at bit-error rates of less than a predefinedthreshold, as indicated in Fig. 3(B.1). In contrast, under badweather conditions, direct links of far nodes are dropped andswitched to other nodes (according to the resource allocationscheme) to keep connectivity to the central node.

Resource allocation in a dynamic cooperative FSO net-work can be optimized for several performance metricsand objectives. Given a number of optical transceivers ineach node (one or two transceivers in our case), the loss co-efficient matrix of FSO links γ, and discrete values of trans-mitted power, many feasible configurations could enablethe kth node (k ∈ f1;…; Ng) to have a bit rate Rk ∈f0; x1; x2;…; xXg at a bit-error rate of less than the thresholdBERk ≤ BERmax. Among these feasible configurations, one ormore could achieve highest network reliability, fairness, andcapacity, and the lowest power and/or bit-error rate. Clearly,the optimized objective functions are conflicted, so each re-source allocation scheme is presented by multiple objectiveoptimization problems (MOOPs). There are several methodsto formulate MOOPs, such as lexicographic (hierarchical),weighted sum, product, and bounded objective functions[19]. Lexicographic optimization is a criterion to optimizeconflicted objectives hierarchically, and it has the ability toachieve resource allocation goals [19,20]. With the lexico-graphicmethod, the objective functions are arranged in orderof importance (optimization levels), and then the objectivefunctions are solved one at a time. On the other hand, theweighted sum scheme optimizes the objectives simultane-ously by transforming the multi-objective function into onescalar objective. There are many forms by which to constructthe weighted sum function; however, the simplest one is thelinear form [19]. In this section, both lexicographic andweight sum optimization are used to formulate two resourceallocation schemes for dynamic cooperative FSO networks.The first scheme considers the optimization of both capacityand fairness (constrained fairness), while the second oneoptimizes the fairness regardless of the capacity.

A. Lexicographic Schemes

In this subsection, two resource allocation schemesare proposed to enhance the performance of dynamic

TABLE INETWORK PARAMETERS

v Node vector �1 × �N � 1��γ Loss coefficient matrix ��N � 1� × �N � 1��G Connection matrix ��N � 1� × �N � 1��Rlkj Bit rate of kth node to node j for lth configurationRlk Bit rate of kth node for lth configurationrl Bit rate vector for lth configuration �1 ×N�R Bit rate matrix �Λ ×N�BERlkj Bit-error rate of kth node to jth node for lth

configurationBERlk Bit-error rate of kth node for lth configurationel Bit-error rate vector for lth configuration �1 ×N�E Bit-error rate matrix �Λ ×N�BERmax Bit-error rate thresholdCl Capacity of the network for lth configurationC Capacity vector (Λ × 1)Cmax Maximum capacityTlij Transmission rate between i and j nodes for lth

configurationTlk Transmission rate of kth node for lth configurationZk Number of transceivers of kth nodePlij Power of i, j link for lth configuration

826 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

Page 6: Fair Resource Allocation Schemes for Cooperative Dynamic

cooperative FSO networks. These schemes are proposed forFSO networks that use two optical transceivers for theinner nodes and one transceiver for the outer nodes, asindicated in Fig. 3(A). The schemes use the concept oflex-max-min fairness, which is widely used in computerand wireless networks to overcome the congestion and lim-ited reliability of the network [21]. Lex-max-min fairnessis a criterion for achieving near equal resource sharingamong N nodes at a relatively high network capacity, i.e.,avoiding inefficient fairness (allocate the lowest bit rate x1for all nodes to achieve the maximum fairness). Lex-max-min fairness is the generalization of ordinarymax-min fair-ness as it searches sequentially for the next maximal, inthe case that two or more solutions have the samemaximalat one level in the space of feasible solutions [21,22].

The first proposed scheme is called lex-max-min con-strained fairness (LMMCF), which aims at enhancing net-work capacity while maintaining fairness between nodes.The second proposed scheme is called lex-max-min fair-ness (LMMF), which aims at enhancing both reliabilityand fairness of the network regardless of the capacity.The LMMCF scheme targets four ordered objective func-tions, which are maximizing network capacity, maximizingbit rate fairness, minimizing power, then finally minimiz-ing bit-error rate. In contrast, the LMMF scheme targetsthree ordered objective functions, which are maximizingbit rate fairness, minimizing power, then finally minimiz-ing bit-error rate.

1) LMMCF: Lexicographic optimization represents theproblem in four levels of optimization based on the prior-ities between the objectives as

Max:l

�Cl �

Xk�N

k�1

Rlk : Cl ∈ C�

Lex-Max-Minl

frl � �Rl1; Rl2;…; RlN� : rl ∈ Rg

Min:l

�Pl �

Xk�N

k�1

Plk : Pl ∈ P�

Lex-Min-Maxl

fel � �BERl1;…;BERlN� : el ∈ Eg

Subject to:

l ∈ f1;2;3;…;Λg; Zk ∈ f1;2g;

Rlk �XNj�0

Rlkj; Rlk ≤ Tlk; Tlk �XNj�0

Tlkj;

BERlk ≤ BERmax; BERlij < BERmax;

Plk �Xj�N

j�0

Plkj; Plkj ∈ f0; y1; y2;…; yYg;

fRlk; Rlkj; Tlk; Tlkjg ∈ f0; x1; x2;…; xXg;k ∈ f1; 2;…; Ng; fi; jg ∈ f0; 1;2;…; Ng; j ≠ i: (14)

This scheme first increases network capacity then pro-ceeds to improve both network reliability and fairness.

The network capacity is the summation of all nodes’ bitrates. Also, the improvement in both reliability and fair-ness are raised by maximizing bit rates of the far nodes.Specifically, at the first optimization level, the LMMCFscheme selects from the feasible Λ configurations the onesthat maximize network capacity. After that, in the secondoptimization level, the scheme searches the previouslyselected configurations for the ones that maximize theminimum bit rate for all nodes.

If there are two or more configurations that have thesame max-min bit rate, the LMMCF scheme proceeds toselect from them the configurations that have the nextmax-min bit rate (sequential max-min optimization) [21].However, if there is more than one configuration withthe same sequential max-min values, the LMMCF selectsfrom them in a third optimization level the configurationsthat have the lowest transmitted power. Finally, if two ormore configurations have the same minimum power, theLMMCF selects from them in a fourth optimization levelthe configuration that minimizes the maximum bit errorrates sequentially (sequential min-max optimization) [22].In addition, only specific discrete values for bit rates, trans-mission rates, and power levels are allowed. In this sense,the bit rates Rlkj associated with node k are selected so thattheir summation Rlk is one of the allowed discrete values.This is also applied to the transmission rate associatedwith each node. Moreover, the average transmitted poweris variable to discrete values. Clearly, in this optimizationproblem, improvement in the bit rate fairness betweendifferent users is restricted by network capacity. Also, animprovement in power efficiency restricts the decreasingof the average error rate.

2) LMMF: Toward achieving the highest network reli-ability and fairness, capacity optimization is omitted whensearching among the feasible configurations. The LMMFscheme selects from the feasible configurations the onesthat maximize minimum bit rate for all nodes in sequen-tial LMM optimization. However, if there is more than onefeasible configuration with the same sequential LMM bitrate, the scheme proceeds to select among these configu-rations the ones that achieve the lowest overall transmit-ted power. Finally, if there is more than one solution,LMMF selects the ones that achieve the sequentialLMM bit-error rates for all nodes. The scheme is pre-sented using lexicographic optimization as three optimiza-tion levels based on the priorities between the objectivesand is described as

Lex-Max-Minl

frl � �Rl1; Rl2;…; RlN� : rl ∈ Rg

Min:l

(Pl �

Xk�N

k�1

Plk : Pl ∈ P

)

Lex-Min-Maxl

fel � �BERl1;…;BERlN� : el ∈ Eg

Subject to:

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 827

Page 7: Fair Resource Allocation Schemes for Cooperative Dynamic

l ∈ f1;2;3;…;Λg; Zk ∈ f1;2g;

Rlk �XNj�0

Rlkj; Rlk ≤ Tlk; Tlk �XNj�0

Tlkj;

BERlk ≤ BERmax; BERlij < BERmax;

Plk �Xj�N

j�0

Plkj; Plkj ∈ f0; y1; y2;…; yYg;

fRlk; Rlkj; Tlk; Tlkjg ∈ f0; x1; x2;…; xXg;k ∈ f1; 2;…; Ng; fi; jg ∈ f0; 1;2;…; Ng; j ≠ i: (15)

Both schemes have the same optimization constraints.However, the capacity achieved by LMMCF is higher thanthat of LMMF, while bit rate fairness achieved in theLMMF case is better than that of LMMCF. This could beinterpreted as follows. In LMMF, the kth node begins toswitch its direct link to another indirect one if its bit rateRk starts to be less than the maximum value xX . This, inturn, decreases the number of direct links in the networkand consequently decreases the capacity. In contrast, forthe LMMCF scheme, the node begins to switch its directlink if its bit rate Rk starts to be less than the minimumallowable value x1. Clearly, this employs the directlinks in order to increase the capacity. In this way,network operators have the ability to select from differentresource allocation schemes that provide different QoSservices.

B. Weighted Sum Schemes

In the lexicographic method, the improvement of the sec-ond-order objective is restricted by the optimized values ofhigher order objectives and, clearly, this could provide thebest formulation of resource allocation in a dynamic co-operative FSO network. However, to compare with otheroptimization formulations, linear weighted sum functionsare constructed in the form U �P

iαi ×Ui, where αi arethe optimization weights, Ui are the optimization objec-tives, and

Piαi � 1. In this method, the normalized perfor-

mance objectives are scaled and combined linearly. Forthe lth configuration, normalized capacity ζ̄l normalizedreliability R̄l, fairness Fl, normalized average transmit-ted power P̄l, and normalized average error rate ξ̄l aregiven as

ζ̄l � Cl∕�N × xX�; 0 ≤ ζ̄l ≤ 1; (16)

R̄l �1N

Xk�N

k�1

δlk; δlk ��1 Rlk > 00 Rlk � 0

; 0 ≤ R̄l ≤ 1;

(17)

Fl � Xk�N

k�1

Rlk

!2

N ×

Xk�N

k�1

R2lk

!; 0 ≤ Fl ≤ 1; (18)

P̄l � Pl∕�⌊�N � n1 � 2 × n2�∕2⌋ × yY�; 0 ≤ P̄l ≤ 1; (19)

ξ̄l �1

Cl × BERmax

Xk�N

k�1

Rlk:Elk; 0 ≤ ξ̄l ≤ 1: (20)

As indicated in Eq. (18), Jain’s index F is used in evalu-ating the bit rate fairness among different N nodes [23].The weighted sum formulation of the proposed tworesource allocation schemes under the same optimizationconstraints is given as

Max:l

Ul � α1 × ζ̄l � α2 × �R̄l � Fl�∕2 − α3 × P̄l − α4 × ξ̄l:

Subject to:

l ∈ f1; 2; 3;…;Λg;

Rlk �XNj�0

Rlkj; Rlk ≤ Tlk; Tlk �XNj�0

Tlkj;

BERlk ≤ BERmax; BERlij < BERmax;

Plk �Xj�N

j�0

Plkj; Plkj ∈ f0; y1; y2;…; yYg;

k ∈ f1;2;…; Ng;Xii�4

ii�1

αii � 1; Zk ∈ f1; 2g;

fRlk; Rlkj; Tlk; Tlkjg ∈ f0; x1; x2;…; xXg;fi; jg ∈ f0; 1; 2; 3;…; Ng; j ≠ i: (21)

Clearly, increasing importance or priority of a perfor-mance object is achieved only by increasing its associatedweight relative to the weights of other objectives in theoptimization equation. Furthermore, using this generalformulation, weighted sum constrained fairness (WSCF)could be formulated by selecting α1 ≫ α2 ≫ α3 ≫ α4. Withthis selection, WSCF aims to give high priority to increas-ing network capacity and less priority to increasing bit ratefairness among different nodes, the same as in LMMCF. Onthe other hand, the weighted sum fairness (WSF) schemecan be formulated by selecting α1 � 0, α2 ≫ α3 ≫ α4, and itclearly performs like the proposed LMMF scheme.

Generally, Eqs. (14), (15), and (21) are classified as integerlinear programming (ILP) multiple objective optimizationproblems (discrete linear MOOPs). These equations canbe solved by using the exhaustive search (ES) method to ob-tain the optimal solution(s). In the ES method, all feasibleconfigurations (β) are first generated using the predefinedoptimization constraints. After that, the configurationsare evaluated for the prioritized or weighted objective func-tions to obtain the optimal solution(s) [24].

The computational complexity of the ES method is of theorder of O�2N�. In small networks (N < 30), ES can be per-formed. However, in large networks, heuristic methods andgenetic algorithms (GAs) are used to obtain sub-optimalsolution(s) with a reasonable number of computationaloperations [25]. The ES method is used to solve the twoILP-MOOPs. However, to overcome the solution complexityfor the proposed schemes, the optimal solution(s) can becomputed offline, then it/they can be stored in a lookup

828 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

Page 8: Fair Resource Allocation Schemes for Cooperative Dynamic

table in an FSO tracking controller to achieve real-timeallocation.

V. SIMULATION AND NUMERICAL RESULTS

In this section, both the LMMCF and LMMF resourceallocation schemes are evaluated and compared to existingschemes to indicate the superior performance of the pro-posed schemes. The evaluations are based on five perfor-mance parameters, which are reliability, capacity, bit ratefairness, average transmitted power, and bit-error rate.These parameters are evaluated under different weatherconditions, which are fog, rain, and snow. Also, some figuresare introduced to indicate the performance differencebetween the WSCF/WSF and LMMCF/LMMF schemesin cooperative dynamic networks.

An NR-OOK format at fixed power level P � −15 dBmand λ � 1550 nm is assumed for most evaluations.However, the impacts of using 780 nm wavelength is dem-onstrated in Figs. 9 and 15. Also, NR-OOK performance iscompared with that of 4-PPM (M � 4), as indicated inFig. 11. Furthermore, the impact of changing transmittedpower level on network performance is shown in Figs. 17and 18.

As shown in Fig. 4, the assumed service area of theconsidered FSO networks is 3 km × 3 km, and nine FSOnodes are assumed to be located uniformly in this area.Moreover, the same homogeneous weather is assumedthroughout the service area. Four topologies are consideredin the evaluations, which are direct link (D-L) [Fig. 2(A)],partial relayed (P-L) [Fig. 2(B)], full relayed (F-L)[Fig. 2(C)], and reconfigurable cooperative [Fig. 3(A)] mod-els. The number of transceivers used for these networks are18, 24, 36, and 22, respectively. All FSO links operate withpredefined six-bit rates, X � 6, at a constant averageoptical power. Table II shows the assigned values of thesimulated FSO network parameters, which are selectedto be in the practical range. From this table, the receiverdiameter is dr � 0.2 m, and the largest FSO link distanceis L � 3700 m. The aperture averaging is calculated asA � �1� 1.33 × �2π∕λ × d2

r∕L��−7∕5 � 0.0038. Clearly, whenusing this value of aperture averaging, the attenuationvariance caused by weak turbulence has small value and

can be neglected [14]. Generally, including attenuationcaused by weak or strong scintillations will not changethe problem formulation of the proposed resource alloca-tion schemes, and it could be directly added to other attenu-ations in the channel loss matrix γ.

Figures 5–9 demonstrate low-visibility atmosphere (fog,low clouds, dust, smoke, etc.) impacts on the considerednetwork performance.

Figure 5 indicates the reliability of the topologies versusvisibility. Clearly, atV ≥ 2.8 km, the nine nodes for all topol-ogies work properly using their direct FSO links. In contrast,at V ≤ 200 m all nodes for all topologies are dropped, i.e.,cannot achieve the minimum bit rate (0.25 Gbps) at a bit-error rate of less than threshold (10−4). Between thesetwo visibility levels, different network topologies have differ-ent performance. At V � 1.4 km, six, three, and two nodesare dropped in D-L, P-L, and LMMF/LMMCF, respectively.Although LMMF and LMMCF have identical performancecurves, in general the reliability of LMMF is better than thatof LMMCF due to its flexibility in link reconfiguration (nocapacity constraint).

Figure 6 shows the capacity of the networks where thecapacity is 9 Gbps for all topologies at V ≥ 3.8 km, and zero

Backbone node

3 km

3km

1km

1 km

0. 5 km

0.5km

(1)(2)

(5)(6)

(7)(8)(9)

(4)

(3)

(0)

FSO node

Fig. 4. Service area of simulated FSO networks.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 40

1

2

3

4

5

6

7

8

9

Dro

pped

nod

es

Visibility (V) in km

P−LD−LF−LLMMCFLMMF

Fig. 5. Network reliability presented in number of dropped nodesversus visibility (V) for D-L, P-L, F-L, LMMCF, and LMMF networks.

TABLE IISIMULATION PARAMETERS

Link Parameters Values

Signal wavelength (λ) 1550 nmDivergence angle (Θ) 2 mm.rad/mDiameter of transmitter (dt) 4 cmDiameter of receiver (dr) 20 cmAverage transmitted signal counts/slot(qt � qs∕γ)

250,000

Average background counts/slot (qb) 50Average transmitted power (P) −15 dBmAverage background noise power −52 dBmDiscrete bit rates (fxX ;…; x2; x1g)in Gbps

1, 3/4, 2/3, 1/2, 1/3, 1/4

Modulation format NR-OOKBER threshold (BERmax) 10−4

Area of FSO network 3 km × 3 kmArea of FSO node cell 1 km × 1 km

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 829

Page 9: Fair Resource Allocation Schemes for Cooperative Dynamic

at V ≤ 200 m. Also, F-L is still the best topology, and thenP-L, since the average transmitted power in these topol-ogies is higher than in other topologies. The performanceof the other three topologies is almost the same; however,the D-L and LMMCF topologies are better than LMMF atcertain V values, as expected. Numerically, at V � 2.2 kmthe capacities of D-L and LMMCF are 6 Gbps (� Cmax),while in LMMF it is 5.5 Gbps. This is due to the maximi-zation capacity optimization in the LMMCF scheme.

Figure 7 explains the bit rate fairness among the nodesversus the visibility for different networks. The maximumfairness is F � 1 at V ≥ 3.4 km for all topologies. Theproposed schemes improve the fairness and outperformthe P-L network, especially at low visibilities. Also, asexpected, at specified visibility values, LMMF performsbetter than LMMCF. Specifically, at V � 1.8 km, thefairness is 1 for both F-L and LMMF, 0.85 for bothLMMCF and P-L, and 0.55 for D-L.

Figure 8 indicates the average bit-error rate of the net-works. Clearly, the LMMF/LMMFC and D-L topologieshave nearly the same performance, and the average bit-error rates are between 10−5 and 10−4 (≤BERmax) for a widerange of visibility. However, both F-L and P-L outperform

the proposed topologies at all visibility values. This is be-cause both LMMF and LMMFC give optimization priorityfor bit rate fairness and capacity, respectively, over the op-timization of bit-error rate. Moreover, the average bit-errorrates for both the F-L and P-L networks are stronglyaffected by atmospheric visibility because there is noresource sharing among different nodes in the networks.

Figure 9 shows the effect of operating optical wavelengthon both reliability and capacity of the LMMCF network atdifferent visibilities. Clearly, the 1550 nm FSO link achieveshigher reliability and capacity than that of the 780 nm FSOlink. Specifically, at V � 1.4 km, the LMMCF network oper-ating at 1550 nm drops two nodes only, while that operatingat 780 nm drops four nodes. In addition, at this visibility, thenetwork achieves capacities of 9 and 7.5 Gbps at 1550 and780 nm, respectively. Generally, shorter wavelengths sufferfrom higher attenuation when propagating in an FSOchannel.

The impacts of rainy weather on the performance of thenetworks considered are demonstrated in Figs. 10 and 11.Figure 10 shows network reliability in terms of number ofdropped nodes versus rainfall rate D. Clearly, for all valuesof rainfall rate and with the same number of implemented

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 40

1

2

3

4

5

6

7

8

9

Visibility (V) in km

Cap

acity

in G

bps

D−LF−LP−LLMMFMMCF

Fig. 6. Capacity versus visibility (V) for D-L, F-L, P-L, LMMF,and LMMCF networks.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 40

0.2

0.4

0.6

0.8

1

Visibility (V) in km

Jain

s F

airn

ess

D−LF−LP−LLMMFLMMCF

Fig. 7. Jain’s bit rate fairness (F) versus visibility (V) for D-L,F-L, P-L, LMMF, and LMMCF networks.

1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 4

10−20

10−15

10−10

10−5

100

Ave

rage

bit

erro

r ra

te

Visibility (V) in km

D−LF−LP−LLMMFMMCF

Fig. 8. Bit-error rate (BER) versus the visibility (V) for D-L, F-L,P-L, LMMF, and LMMCF networks.

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 60

1

2

3

4

5

6

7

8

9

10

Dro

pped n

odes

Visibility (V) in km0 1 2 3 4 5

0

2

4

6

8

10

0 1 2 3 4 50

2

4

6

8

10

Capaci

ty in

Gpbs

LMMCF−1550

LMMCF−780

Reliability

Capacity

Fig. 9. Reliability and capacity of the LMMF scheme at two differ-ent operating wavelengths, 1550 and 780 nm, versus visibility (V).

830 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

Page 10: Fair Resource Allocation Schemes for Cooperative Dynamic

optical transceivers, both the LMMF and LMMCF schemeshave higher reliability than the P-L scheme. Numerically,at D � 10 mm∕h, five, two, and one node(s) are droppedin the D-L, P-L, and LMMF/LMMCF networks, respectively.In addition,Fig. 11 explains themodulation impact onLMMFscheme performance, where there are remarkable improve-ments in network reliability and capacity performance for4-PPM (2 bits/frame) over NR-OOK at the same averagetransmitted power and background noise. Numerically, atD � 30 mm∕h, the network drops three and five nodes for4-PMM and NR-OOK, respectively. Also, at D � 10 mm∕h,the capacities are 4 Gbps and 5 Gbps for 4-PMM andNR-OOK, respectively. Clearly, PPM has higher immunityagainst severe weather conditions over the OOK format,which enables the direct link to survive and hence increasenetwork capacity. Reference [11] shows the rest of the resultsfor rainy channels for the networks considered.

Figures 12–15 demonstrate wet snow impacts on the per-formance of the networks considered. Figure 12 explainsthe reliability versus snow fall rateSw. Specifically, at Sw �10 mm∕h, eight, seven, and six nodes are dropped in theD-L/P-L, LMMF/LMMCF, and F-L models, respectively.

Moreover, on the one hand, as indicated in Fig. 13, theF-L network maintains C � 9 Gbps until Sw � 2 mm∕h,

and with a severe atmosphere, Sw � 20 mm∕h, F-L stillmaintains about C � 1 Gbps. However, as expected, atdifferent Sw values, the performance of LMMCF is betterthan that of the LMMF scheme. On the other hand,Fig. 14 shows the bit rate fairness among the nodesversus Sw, where the P-L network maintains F � 1 until

101

102

0

1

2

3

4

5

6

7

8

9

Rain fall rate (D) in mm/h

Dro

pped

nod

es

D−LP−LF−LLMMCFLMMF

Fig. 10. Reliability versus the rain fall rate (D) for D-L, F-L, P-L,LMMF, and LMMCF networks.

5 30 60 130 1802 200

1

2

3

4

5

6

7

8

9

10

Cap

acity

in G

pbs

Rain fall rate (D) in mm/h10

110

20

2

4

6

8

10

101

102

0

2

4

6

8

10

Dro

pped

nod

es

LMMF−OOK

LMMF−4PPM

Reliability

Capacity

Fig. 11. Reliability and capacity of LMMF scheme for OOK and4-PPM modulation versus rain fall rate (D).

1 2 5 10 200

1

2

3

4

5

6

7

8

9

Wet snow fall rate (Sw) in mm/h

Dro

pped

Nod

es

D−LP−LF−LLMMCFLMMF

Fig. 12. Reliability versus wet snow fall rate (Sw) for D-L, F-L,P-L, LMMF, and LMMCF networks.

1 105 2020

1

2

3

4

5

6

7

8

9

Wet snow fall rate (Sw) in mm/h

Cap

acity

in G

bps

D−LF−LP−LLMMF

Fig. 13. Capacity versus wet snow fall rate (Sw) for D-L, F-L, P-L,LMMF, and LMMCF networks.

1 2 5 10 200

0.2

0.4

0.6

0.8

1

Wet snow fall rate (Sw) in mm/h

Jain

Fai

rnes

s (F

)

D−LP−LF−LLMMCFLMMF

Fig. 14. Fairness (F) versus wet snow fall rate (Sw) for D-L, F-L,P-L, LMMF, and LMMCF networks.

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 831

Page 11: Fair Resource Allocation Schemes for Cooperative Dynamic

Sw � 1 mm∕h and outperforms the proposed schemes,while, with a worse atmosphere of Sw ≥ 1.5 mm∕h, bothLMMF then LMMCF achieve better fairness performancethan the P-L topology. In addition, as indicated in Fig. 15for the LMMCF scheme, insignificant enhancementsare achieved in both network reliability and capacity forthe 780 nm FSO link over the 1550 nm FSO link.Particularly, the first node is dropped at Sw � 1.43 mm∕hand Sw � 1.47 mm∕h for the 1550 and 780 nm FSO links,respectively.

The impacts of dry snow weather on the performanceof the networks considered are demonstrated in Fig. 16.Figure 16 shows that, at Sd � 2.5 mm∕h, the number ofdropped nodes for the D-L/P-L, LMMF/LMMCF, and F-Lnetworks are eight, five, and three nodes, respectively.The details of impacts for dry snow weather on the perfor-mance of the networks considered are demonstrated in [10].

Generally, the previous results show that the LMMCF/LMMF schemes offer better performance in reliability andfairness in different weather conditions at lower networkcost. Further improvement could be achieved by using appro-priate wavelength andmodulation schemes. Also, the impactof dry snow is the strongest over both wet snow and rain atthe same fall rate. Precisely, the considered networks are

completely dropped at Sd � 9 mm∕h, Sw � 20 mm∕h, andD � 180 mm∕h as shown in Figs. 16, 12, and 10, respectively.

Figures 17 and 18 compare the performance of theLMMCF and LMMF schemes, respectively, under bothfixed and adaptive transmitted powers. For the fixed powerscheme, the transmitted power is P ∈ f0;−15g dBm, whilefor the adaptive power scheme, the transmitted power isP ∈ f0;−18;−17;−16.5;−16;−15.5;−15g dBm. Clearly, asindicated in these figures, adaptive power schemes achievemuch higher power efficiency than that of the fixed ones inboth the LMMCF and LMMF schemes. Clearly, this comeswith the price of increasing the average bit-error rate in thenetwork. Numerically, for the LMMCF scheme with foggyweather as indicated in Fig. 17, at V � 2 km, the averagetransmitted powers for cases of adaptive power and fixedpower are −8 and −5 dBm, respectively. Also, at this visibil-ity value, the average error rates for the adaptive powerand fixed power schemes are 0.5 × 10−4 and 1−5, respec-tively. Figure 18 shows the same behavior for the LMMFscheme under rainy weather conditions.

Figures 19 and 20 indicate the performance differencebetween the lexicographic and weighted sum schemes,and in our case all schemes achieved the same reliability

3 5 7 15 200

1

2

3

4

5

6

7

8

9

10

Cap

acity

in G

pbs

Wet snow fall rate (Sw) in mm/h3 5 7 15 20

0

2

4

6

8

10

100

101

0

2

4

6

8

10

Dro

pped

nod

es

LMMCF−1550LMMCF−780

Reliability

Capacity

Fig. 15. Reliability and capacity of LMMCF scheme at two differ-ent wavelengths, 1550 and 780 nm, versus wet snow fall rate (Sw).

1 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.50

1

2

3

4

5

6

7

8

9

Dry snow fall rate (Sd) in mm/h

Dro

pped

Nod

es

D−LP−LF−LLMMCF

Fig. 16. Reliability versus dry snow fall rate (Sd) for D-L, F-L,P-L, LMMF, and LMMCF networks.

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3−20

−17

−14

−11

−8

−5

−2

0

Visibility (V) in Km

Ave

rage

tran

smitt

ed p

ower

(dB

m)

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3

0.5

1

1.5

2

2.5

x 10−4

Ave

rage

err

or r

ate

0.5

1

1.5

2

2.5

x 10−4

Fixed Power LevelAdaptive Power Level

Error rate

Power

Fig. 17. Average transmitted power and average error rate forfixed and adaptive power LMMCF schemes in a foggy channel.

3 4 7 10 13 16 21 31 135−20

−17

−14

−11

−8

−5

−2

0

Rain fall rate (D) in mm/h

Ave

rage

tran

smitt

ed p

ower

(dB

m)

40

0.5

1

1.5

2

2.5

x 10−4

Ave

rage

err

or r

ate

0.5

1

1.5

2

2.5

x 10−4

Fixed Power Level

Adaptive Power LevelPower

Error rate

Fig. 18. Average transmitted power and average error rate forfixed and adaptive power LMMF schemes in a rainy channel.

832 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.

Page 12: Fair Resource Allocation Schemes for Cooperative Dynamic

performance for a foggy channel, as shown in Fig. 5. Theformulations of the WSCF and WSF schemes are selectedto be comparable to LMMCF and LMMF, respectively,where in WSCF the selected weights are α1 � 0.5,α2 � 0.3, and α4 � 0.2, while in WSF the selected weightsare α1 � 0, α2 � 0.8, and α4 � 0.2 (α3 � 0 for fixed powerlinks). As indicated by Fig. 19, LMMCF always achievesthe highest capacity over the entire visibility range.Also, WSCF achieves higher capacity than that of theLMMF and WSF schemes. On the other hand, Fig. 20 in-dicates the superior performance of both LMMF and WSFin terms of bit rate fairness compared to other schemes.However, the performance of the WSCF and WSF schemesis strongly affected by the selected value of the optimiza-tion weights. Clearly, as indicated in Fig. 20, an increaseof weight α1 results in significant enhancement in networkcapacity associated with an inherent decrease in the bitrate fairness among the different nodes.

Last, it is worth noting that both reliability and bit ratefairness of the dynamic cooperative FSO network could besignificantly enhanced by allocating more optical trans-ceivers to the nodes near the backbone node. Clearly, thesenodes have less distance to the central node and theirdirect FSO links have better qualities in severe weather

conditions. However, the optimal transceiver distribution(s)[optimal transceiver placement(s)] Zk is out of the scope ofthis paper and could be investigated in future researchwork to obtain the optimal transceiver distribution inreconfigurable cooperative FSO networks.

VI. CONCLUSION

Two resource allocation schemes have been proposed toimprove both the reliability and fairness of dynamic co-operative FSO networks. One of these schemes enhancesthe network capacity over the bit rate fairness and theother does the opposite. The proposed schemes have as-sumed that the outer nodes have one optical transceiver,while the inner nodes are equipped with two transceivers.Each scheme is formulated as a discrete linear multi-objec-tive optimization problem (ILP-MOP). Both schemes areevaluated and compared to relayed networks (partialand full relayed) under different weather conditions, e.g.,fog, rain, and snow. Using the same number of opticaltransceivers, our results reveal that the proposed schemesachieve better reliability and bit rate fairness over partialrelayed networks at all weather conditions. In addition, theresults indicate that the impact of dry snow droplets onnetwork performance is higher than both wet snow andrain droplets at the same fall rate. As future research,the network performance could be evaluated for differentallocations of additional optical transceivers on differentnodes, and the optimal distribution of these transceiverscould be obtained.

ACKNOWLEDGMENT

This work is supported by the Ministry of HigherEducation (MoHE) of Egypt and is funded by an STDFgrant #12670, Science and Technology DevelopmentFund, Egypt.

REFERENCES

[1] I. I. Kim, B. McArthur, and E. J. Korevaar, “Comparison oflaser beam propagation at 785 nm and 1550 nm in fog andhaze for optical wireless communications,” Proc. SPIE,vol. 4214, pp. 26–37, 2001.

[2] H. H. Refai, J. J. Sluss, H. H. Refai, and M. Atiquzzaman,“Comparative study of the performance of analog fiber opticlinks versus free-space optical links,” Opt. Eng., vol. 45, no. 2,025003, 2006.

[3] S. Bloom, E. Korevaar, J. Schuster, and H. Willebrand,“Understanding the performance of free-space optics[Invited],” J. Opt. Netw., vol. 2, no. 6, pp. 178–200, 2003.

[4] “Prediction methods required for the design of terrestrialfree-space optical links,” ITU-R Recommendation P.1814,2007.

[5] A. Vavoulas, H. G. Sandalidis, and D. Varoutas, “Weathereffects on FSO network connectivity,” J. Opt. Commun.Netw., vol. 4, no. 10, pp. 734–740, 2012.

[6] M. A. Kashani, M. Safari, andM. Uysal, “Optimal relay place-ment and diversity analysis of relay-assisted free-space opti-cal communication systems,” J. Opt. Commun. Netw., vol. 5,no. 1, pp. 37–47, 2013.

1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.60.4

0.5

0.6

0.7

0.8

0.9

1

Visibility (V) in Km

Fai

rnes

s

1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.64

4.5

5

5.5

6

6.5

7

7.5

8

8.5

9

4

4.5

5

5.5

6

6.5

7

7.5

8

8.5

9

Cap

acity

in G

bps

LMMF WSF LMMCF WSCF

Fairness

Capacity

Fig. 19. Capacity and fairness versus the visibility (V) for lexico-graphic and weighted sum schemes.

5

6

7

8

9

10

Visibility (V) in Km

Ca

pa

city

in G

pb

s

2 2.2 2.4 2.6 2.8 30.75

0.8

0.85

0.9

0.95

1

Fa

irn

ess

α1=0.33, α

2=0.33, α

3=0.0, α

4=0.33

α1=0.5, α

2=0.5, α

3=0.0, α

4=0.0

α1=0.0, α

2=0.8, α

3=0.0, α

4=0.2

α1=0.5, α

2=0.3, α

3=0.0, α

4=0.2

Fairness

Capacity

Fig. 20. Capacity and fairness versus visibility (V) for different αiicoefficients for weighted sum schemes.

Ghazy et al. VOL. 8, NO. 11/NOVEMBER 2016/J. OPT. COMMUN. NETW. 833

Page 13: Fair Resource Allocation Schemes for Cooperative Dynamic

[7] Z. Hu, P. Verma, and J. Sluss, Jr., “Improved reliability of free-space optical mesh networks through topology design,” J. Opt.Netw., vol. 7, no. 5, pp. 436–448, 2008.

[8] S. D. Milner, T.-H. Ho, I. I. Smolyaninov, S. Trisno, and C. C.Davis, “Free-space optical wireless links with topology con-trol,” Proc. SPIE, vol. 4821, pp. 175–180, 2002.

[9] A. S. Ghazy, H. A. I. Selmy, and H. M. H. Shalaby, “Faircooperative resource allocation schemes for foggy free spaceoptical network,” in Int. Conf. on Transparent OpticalNetworks (ICTON), 2015, vol. 2, pp. 178–200.

[10] A. S. Ghazy, H. A. I. Selmy, and H. M. H. Shalaby, “Reliable-fair resource allocation schemes for snowy cooperative freespace optical networks,” in Asia Communications andPhotonics Conf. (ACP), 2015, vol. 3, pp. 278–300.

[11] A. S. Ghazy, H. A. I. Selmy, and H. M. H. Shalaby, “Proposedresource allocation schemes for rainy free space optical net-work,” in Proc. 4th Int. Conf. on Photonics, Optics and LaserTechnology, 2016, pp. 74–81.

[12] R. M. Gagliardi and S. Karp, Optical Communications,2nd ed. New York: Wiley, 1995.

[13] J. H. Churnside, “Aperture averaging of optical scintillationsin the turbulent atmosphere,” Appl. Opt., vol. 30, no. 15,pp. 1982–1994, 1991.

[14] L. C. Andrews, R. L. Phillips, and C. Y. Hopen, Laser BeamScintillation With Applications. Bellingham, WA: SPIE,2001, vol. 99.

[15] L. C. Andrews and R. L. Phillips, Laser Beam PropagationThrough Random Media. Bellingham, WA: SPIE, 2005,vol. 52.

[16] M. A. Khalighi and M. Uysal, “Survey on free space opticalcommunication: A communication theory perspective,”IEEE Commun. Surv. Tutorials, vol. 16, no. 4, pp. 2231–2258, 2014.

[17] O. Barsimantov and V. V. Nikulin, “Adaptive optimization of afree space laser communication system under dynamic linkattenuation,” J. Opt. Commun. Netw., vol. 3, no. 3, pp. 215–222, 2011.

[18] P. T. Dat, A. Bekkali, K. Kazaura, K. Wakamori, and M.Matsumoto, “A universal platform for ubiquitous wireless com-munications using radio over FSO system,” J. LightwaveTechnol., vol. 28, no. 16, pp. 2258–2267, 2010.

[19] R. T. Marler and J. S. Arora, “Survey of multi-objective opti-mization methods for engineering,” Struct. Multidiscip.Optim., vol. 26, no. 6, pp. 369–395, 2004.

[20] H. Isermann, “Linear lexicographic optimization,” ORSpectrum, vol. 4, no. 4, pp. 223–228, 1982.

[21] W. Ogryczak and T. Sliwinski, “Lexicographic max-min opti-mization for efficient and fair bandwidth allocation,” in Int.Network Optimization Conf. (INOC), 2007.

[22] W. Ogryczak and T. Sliwinski, “On direct methods for lexico-graphic min-max optimization,” in Computational Scienceand Its Applications (ICCSA), 2006, pp. 802–811.

[23] R. Jain, D. Chiu, and W. Hawe, “A quantitative measure offairness and discrimination for resource allocation in sharedcomputer systems,” arXiv:cs/9809099, 1998.

[24] C. Paar and J. Pelzl, Understanding Cryptography: ATextbook for Students and Practitioners. New York: Springer,2009.

[25] D. E. Golberg, Genetic Algorithms in Search, Optimization,and Machine Learning. Boston: Addison-Wesley, 1989.

Abdallah S. Ghazy was born in Giza,Egypt, in 1983. He received the B.S. degreefromAzher University, Egypt, in 2007 in elec-trical engineering. He worked in voice-datasystems in Telecom-Egypt and AVAYA-Egypt for six years, from 2008 to 2013.In 2012, he joined the Electronics andElectrical Communications EngineeringDepartment, Azher University, Egypt. Heis currently on leave from Azher University

and is working toward the M.S. degree at Egypt-Japan Universityof Science and Technology, Alexandria, Egypt. His research in-terests include optical communication, wireless communication,and heterogeneous communication networks.

Hossam A. I. Selmy was born in Giza,Egypt, in 1979. He received the B.S. andM.S. degrees from Cairo University, Egypt, in2001 and 2007, respectively, and the Ph.D.degree from Egypt-Japan University forScience and Technology, Egypt, in 2013, allin electrical engineering. He is currently anAssistant Professor at the National Instituteof Laser Enhanced Science (NILES), CairoUniversity. His research interests include

advanced modulation and multiple access schemes for optical fibercommunication and next-generation wireless access networks.

Hossam M. H. Shalaby (S’83–M’91–SM’99)was born in Giza, Egypt, in 1961. He re-ceived the B.S. and M.S. degrees fromAlexandria University, Alexandria, Egypt,in 1983 and 1986, respectively, and thePh.D. degree from the University ofMaryland at College Park in 1991, all inelectrical engineering. In 1991, he joinedthe Electrical Engineering Department,Alexandria University, and was promoted

to Professor in 2001. From September 2010 to August 2016,he was the Chair of the Department of Electronics andCommunications Engineering, Egypt-Japan University of Scienceand Technology (E-JUST), Alexandria, Egypt. From September1996 to January 1998, he was with the Electrical and ComputerEngineering Department, International Islamic UniversityMalaysia, and from February 1998 to February 2001, he was withthe School of Electrical and Electronic Engineering, NanyangTechnological University, Singapore. His research interests in-clude optical communication, silicon photonics, optical space-division multiplexing, optical CDMA, and information theory.He is a senior member of the IEEE Photonics Society and TheOptical Society (OSA).

834 J. OPT. COMMUN. NETW./VOL. 8, NO. 11/NOVEMBER 2016 Ghazy et al.