Skip to main content
Erschienen in: EURASIP Journal on Wireless Communications and Networking 1/2010

Open Access 01.12.2010 | Research Article

Efficient Uplink Modeling for Dynamic System-Level Simulations of Cellular and Mobile Networks

verfasst von: Ingo Viering, Andreas Lobinger, Szymon Stefanski

Erschienen in: EURASIP Journal on Wireless Communications and Networking | Ausgabe 1/2010

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

A novel theoretical framework for uplink simulations is proposed. It allows investigations which have to cover a very long (real-) time and which at the same time require a certain level of accuracy in terms of radio resource management, quality of service, and mobility. This is of particular importance for simulations of self-organizing networks. For this purpose, conventional system level simulators are not suitable due to slow simulation speeds far beyond real-time. Simpler, snapshot-based tools are lacking the aforementioned accuracy. The runtime improvements are achieved by deriving abstract theoretical models for the MAC layer behavior. The focus in this work is long term evolution, and the most important uplink effects such as fluctuating interference, power control, power limitation, adaptive transmission bandwidth, and control channel limitations are considered. Limitations of the abstract models will be discussed as well. Exemplary results are given at the end to demonstrate the capability of the derived framework.

1. Introduction

The requirements for simulation tools are changing with the introduction of novel advanced methods. In particular, investigation of self-organizing networks (SONs) [15] have to cover extremely long time intervals; however, they require a sufficient level of accuracy in terms of radio resource management (RRM), quality of service (QoS), and mobility at the same time. For instance, self-optimization of the downtilt angle [6] is a process which may cover at least several days, since the network has to make sure that meaningful statistics on user locations and signal strengths have been collected. Furthermore, there are certainly interactions and collisions between SON and RRM, so that RRM cannot be entirely excluded from the simulations. For instance, if the downtilt angle is changed too fast, RRM measurements might get confused leading to an unstable system. Similar things hold for other SON use cases such as load balancing [7], mobility robustness optimization, and automatic neighbor relation [5].
Typical system-level simulations [8] have a very exact implementation of RRM and QoS by explicitly modeling all the fast decisions, typically on a millisecond time scale or even below, for example [9]. This ends up in a very large simulation runtime, far beyond real-time. Simulating several hours, days, or even more is impossible with this class of simulators. Those simulators are used to make accurate performance evaluations given a fixed parameter configuration according to specified reference scenarios.
Alternatively, the use of light, snapshot-based tools is quite popular [10, 11]. Those allow for a rapid collection of network statistics. However, accuracy of RRM and QoS is lost to a wide extent. In particular, handover effects such as hysteresis and time to trigger. can not be modeled without having a true time axis implemented. Furthermore, traffic characteristics are poorly reflected, for example, the fact that users at the cell edge require much more resources than close users in many cases. It is also more than critical to investigate convergence behavior of dynamic SON loops without a real-time axis and without real mobility. Those simulators are used for network planning or for coarse studies to understand the interrelations of new features, for example, heterogeneous networks [12].
In this work we will present the theoretical framework for a new class of simulators which is capable of making very long SON simulations with the necessary level of accuracy. It can be understood as a smart extension of snapshot-based tools with a time axis and with abstract, semianalytical models of RRM and QoS. It allows self-tuning of parameters during the simulations (which is a typical SON aspect) rather than using a fixed parameter configuration for every simulation. We are certainly not reaching the accuracy of full system-level simulations; however, this is not needed in many cases. For the downlink this work has already been started in [13]. Unfortunately the uplink shows a lot of fundamental differences compared with the downlink which complicates mattersin the following way.
(i)
Every terminal has its own individual power budget.
 
(ii)
The uplink typically has a power control (due to near/far problem).
 
(iii)
The intercell interference is heavily fluctuating.
 
(iv)
Control channel limitations are more critical.
 
(v)
The access scheme might be different so that the scheduling strategies are different.
 
Those aspects will be addressed in this work based on the principles introduced in [13]. Although the focus of this work is on the introduction of the simulation framework, we will also give some calibration results as well as some first SON results. The derivations are based on the 3GPP standard long-term evolution (LTE) [14]. However the principles can be applied to other systems such as HSPA and WiMAX as well.
We will start with definitions of the LTE uplink, the uplink power control, and the uplink SINR. In Section 3 we will discuss the scheduling strategies. We will consider different resource fair strategies, throughput fair strategies and QoS strategies targeting a certain bit rate. All derivations are done under the assumption of anadaptive transmission bandwidth scheduler. Performance metrics are introduced in Section 4, in particular, dissatisfaction levels due to overload, power limitation, and control channel limitation. Results with the new framework are given in Section 5, and Section 6 concludes this work. In the appendices important and interesting properties of fairness in the uplink in comparison to downlink fairness are discussed.

2. Definitions

We will discuss the LTE uplink, which is a Single Carrier FDMA system. [14]. The whole system bandwidth is divided into https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq1_HTML.gif subbands which are called physical resource blocks (PRBs). In every transmission time interval (TTI) a user can be assigned a subset of those https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq2_HTML.gif PRBs which, however, have to be adjacent. The user will spread the symbols to transmit over this group of PRBs. Note that this so-called single carrier constraint is different to the OFDMA downlink.
Due to the single carrier constraint a frequency selective scheduler for the LTE uplink may have a packing problem ("Tetris" problem), that is, it might not be able to fill the entire bandwidth in some cases. The more multiuser diversity the scheduler aims to exploit, the larger will be the packing problem. In this work we neglect those cutaways, that is, we assume that the scheduler can fill the entire bandwidth. Note that it is very easy to construct such a scheduler, but the frequency-selective multi-user gain will be poor.
Random variables will be written in bold letters, for example, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq3_HTML.gif or https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq4_HTML.gif . It is very important for this work to distinguish between random and deterministic variables. All variables refer to linear values, except the first equations (1) to (4) that make use of the dB domain. For the sake of better notation we are using the same symbols nevertheless.

2.1. General Definitions

We are assuming a network given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq5_HTML.gif users https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq6_HTML.gif located at the coordinates https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq7_HTML.gif , and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq8_HTML.gif cells https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq9_HTML.gif . All propagation effects (comprising pathloss, antenna patterns, and shadowing) between position https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq10_HTML.gif and cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq11_HTML.gif are summarized in the propagation maps https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq12_HTML.gif . Details on the included propagation effects are found in [13]. Note that the propagation maps are deterministic for our investigations even if the shadowing has been generated randomly. Fast Fading is not considered in this work. https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq13_HTML.gif is the thermal noise on a single PRB.
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq14_HTML.gif is the downtilt angle of cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq15_HTML.gif . We assume that this is the only propagation parameter which can be dynamically influenced, all others are either given by the environment (e.g., pathloss exponent, shadowing) or are configured statically (e.g., antenna height, azimuth orientation) and are therefore omitted. Please note that downtilt optimization is an important SON use case, and hence we leave the downtilt angle in the equations although we do not present results on that.
Furthermore, every cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq16_HTML.gif can adjust individual power control settings given by the parameters https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq17_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq18_HTML.gif according to [15]. We assume that user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq19_HTML.gif is served by cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq20_HTML.gif , where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq21_HTML.gif is theconnection function, and every user is connected exactly to a single cell. In this work, we assume that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq22_HTML.gif is given by the best serving cell on downlink, that is, every user is connected to the strongest cell. This is a typical case; however we could in principle also optimize the connection function with the equations given in this work.
The number of users in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq23_HTML.gif is abbreviated by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq24_HTML.gif , and the set of users connected to cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq25_HTML.gif is abbreviated by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq26_HTML.gif .

2.2. Power Control

Uplink Power Control is typically given by the equation (cf. [15], neglecting the closed loop terms)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ1_HTML.gif
(1)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq27_HTML.gif is the total transmit power of user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq28_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq29_HTML.gif is the maximum transmit power, and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq30_HTML.gif is the number of PRBs allocated to user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq31_HTML.gif . In the following we will use the transmit power per PRB https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq32_HTML.gif instead of the total transmit power https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq33_HTML.gif . Furthermore, we assume that the scheduler at the serving cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq34_HTML.gif is smart enough that it will not drive users into power limitation through the choice of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq35_HTML.gif , that is it will limit the number of PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq36_HTML.gif such that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq37_HTML.gif operator does not expire (the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq38_HTML.gif operator can only expire for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq39_HTML.gif ). This behavior will be elaborated later on in Section 3.2. In this case we can define the transmit power per PRB (actually power spectral density) as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ2_HTML.gif
(2)

2.3. Signal-to-Noise and Interference Ratio

With this definition, we can write the received power of user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq40_HTML.gif at its serving cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq41_HTML.gif as (we are omitting the superscript https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq42_HTML.gif for the following variables although we keep on using spectral densities/power per PRB)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ3_HTML.gif
(3)
Similarly, we define the interference produced by user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq43_HTML.gif at any other cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq44_HTML.gif as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ4_HTML.gif
(4)
Note that this interference is only produced if user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq45_HTML.gif is scheduled by its serving cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq46_HTML.gif at the time and PRB of interest. Let us define the random variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq47_HTML.gif which specifies the user which is scheduled by cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq48_HTML.gif at a particular time and a particular PRB. We call the probability that cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq49_HTML.gif schedules user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq50_HTML.gif the scheduling probabilities https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq51_HTML.gif . We assume that the scheduling probabilities are identically distributed over time and frequency but not independently. Correlations and further details of the random variables https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq52_HTML.gif will be discussed later on. As a consequence, the interference produced from cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq53_HTML.gif to a target cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq54_HTML.gif is also a random variable:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ5_HTML.gif
(5)
Furthermore the SINR for user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq55_HTML.gif also gets a random variable (although we ignore fast fading at all):
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ6_HTML.gif
(6)
Note that whereas we have used power values in dB so far, any power and SINR variables in this and the following equations are linear values (using the same symbols). In the following we will look at the average of this random SINR (still on a per user basis):
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ7_HTML.gif
(7)
Let us make some important observations.
(i)
The received power https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq56_HTML.gif is not a random variable.
 
(ii)
The last expectation of (7) does not depend on user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq57_HTML.gif , only on the cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq58_HTML.gif , that is, it is the same for all other users connected to cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq59_HTML.gif .
 
(iii)
It is interesting to see that the more the interference https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq60_HTML.gif fluctuates, the smaller gets the average SINR. This is easily derived from Jensen's inequality ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq61_HTML.gif is a convex function).
 
Note that the random variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq62_HTML.gif is actually a deterministic function of the random variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq63_HTML.gif (cf. (5)),that is, the interference is determined as soon as the scheduler has selected a user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq64_HTML.gif .

2.4. Evaluation of the Expectation

Even if we already knew the scheduling probabilities https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq65_HTML.gif , the expectation would be very inconvenient to evaluate. In this section, we assume that the scheduling probabilities are well known (we will discuss later on how to calculate them), and we will focus on the evaluation of the expectation in the average SINR expression (7). We have observed that this expectation is cell specific and does not depend on the user, so we have replaced https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq66_HTML.gif directly by cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq67_HTML.gif :
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ8_HTML.gif
(8)
Obviously, this expectation is multidimensional, since https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq68_HTML.gif different (independent) random variables https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq69_HTML.gif 's are involved. We can give a closed-form expression:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ9_HTML.gif
(9)
Please note that cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq70_HTML.gif does not contribute to the interference on itself. However, for the sake of better illustration we have left the corresponding sum in the equation. Unfortunately, the nested sum can hardly be evaluated numerically. For instance, in a typical scenario [16] with 57 cells and 10 users per cell we would have https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq71_HTML.gif addends. Unfortunately, due to the nonlinearity of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq72_HTML.gif function, there is no way to separate the random variables and thereby the nested sums. Restricting the interference impact to only close neighbors (e.g., first and second ring around a cell) reduces the problem a bit; however it is still hardly feasible. Note that we have used the abbreviation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq73_HTML.gif which is the set of users connected to cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq74_HTML.gif .
A practical solution is a Monte Carlo integration. We generate a large number https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq75_HTML.gif of random https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq76_HTML.gif - tuples https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq77_HTML.gif with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq78_HTML.gif containing samples of the random variables https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq79_HTML.gif . As long as the number of samples https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq80_HTML.gif is sufficiently large, we can get a good approximation of the expectation by
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ10_HTML.gif
(10)
Our investigations have shown that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq81_HTML.gif gives stable results and is still feasible from a complexity point of view. Note that for the Monte Carlo approach the generation of the random https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq82_HTML.gif -tuples certainly must follow the scheduling probabilities https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq83_HTML.gif . Accuracy can be increased by combining the two approaches: the first ring of interfering cells can be exactly evaluated whereas the rest of the cells is considered by the Monte Carlo approach. In this paper we have only used the Monte-Carlo approach.

2.5. Rate Function

Using the previously derived https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq84_HTML.gif (per PRB) we define a rate function https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq85_HTML.gif to be the data rate which a user can achieve on a single PRB with average https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq86_HTML.gif using an appropriate modulation and coding scheme. In the simplest case we could use Shannon's capacity equation or an extension thereof. In this work, we will follow a more realistic approach using link level results. We are using an abstract model presented in [17] which has been shown to be very close to real simulations using the Turbo codes defined in 3GPP [14]. The LTE uplink overhead through reference signals has been taken into account. Figure 1 shows the employed rate function including the Shannon reference with and without considering the LTE overhead.
Note that the Shannon bounds inherently assume a perfect selection of modulation and coding schemes. However in the uplink, due to fluctuating interference, this selection can not be perfect by definition, even not in static channel conditions. Furthermore imperfect channel estimation will also degrade the performance. The consequence is a loss of some dBs. On the other hand, the base stations typically have 2 receive antennas, which is also not considered in the Shannon bounds which will lead to a gain in the range of 3 dB. Furthermore, frequency selective scheduling (e.g., though proportional fair scheduling) will lead to multi-user diversity gain [18, 19].
In this work we will assume that those effects will compensate each other such that the rate function used here (red solid curve) is rather close to the Shannon bound considering the overhead through cyclic prefix and reference signals. Later on in Section 5.2 we will see that this assumption leads to a good agreement with existing simulation results.

3. Scheduling Probabilities

Let us now have a closer look at the scheduling probabilities https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq87_HTML.gif . We will consider several scheduler strategies. Note that the random variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq88_HTML.gif is discrete; it can adopt values https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq89_HTML.gif with the probability https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq90_HTML.gif . For mathematical correctness, we need to define a kind of idle value, for example, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq91_HTML.gif , with nonzero probability https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq92_HTML.gif which represents the case that no user is scheduled in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq93_HTML.gif (at the considered time and frequency, that is, a PRB is left empty). All other values have the probability https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq94_HTML.gif . With these definitions, we can write (just for comprehension)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ11_HTML.gif
(11)

3.1. General Expression

Let us define the average number of PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq95_HTML.gif which is allocated to user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq96_HTML.gif . Note that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq97_HTML.gif . Given all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq98_HTML.gif 's in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq99_HTML.gif , we can write the scheduling probabilities as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ12_HTML.gif
(12)
We observe that the scheduling probabilities depend purely on the average number of assigned PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq100_HTML.gif 's. Hence, we will investigate those elaborately in the following sections. We will be looking at individual cells; we assume that cells in general behave independently, that is, the random variables https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq101_HTML.gif 's are mutually independent, too.

3.2. Adaptive Transmission Bandwidth

The key difference compared with the downlink is the fact that every user has an individual power budget in the uplink. So we can shift PRBs from one user to another, but not power. As a direct consequence, the maximum number of PRBs which can be given to a user without driving it into power limitation depends on the difference between transmit power per PRB https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq102_HTML.gif (given by (2)) and the maximum transmit power https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq103_HTML.gif which is typically called power headroom:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ13_HTML.gif
(13)
An uplink scheduler should never assign a user more PRBs than this limit https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq104_HTML.gif . Otherwise, looking at the original power control equation (1), we observe that the users would have to spread the same power over the assigned PRBs instead of increasing the power with every assigned PRB (the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq105_HTML.gif operator in the PC equation (1) expires). This results in an SINR loss which would eat up at least part of the bandwidth gain. Furthermore, other (non-power-limited) users can make much better use of the bandwidth. Finally, spreading the maximum power over several PRBs would increase the dynamic range problems. Note that for the PC equation per PRB (2) we have already inherently assumed that the scheduler does not exceed the aforementioned limit. This behavior is typically called adaptive transmission bandwidth [ 20 ].
Obviously this limits the maximum average number of PRBs as well, since every user can be scheduled at maximum in every time slot, hence we have
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ14_HTML.gif
(14)

3.3. Strict Resource Fair

The straightforward definition of the resource fair scheduler would be that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq106_HTML.gif users in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq107_HTML.gif share the available resources, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq108_HTML.gif . However, this may violate the power limitation of the UEs in (14). If we require resource fairness, nevertheless, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq109_HTML.gif should be the same for all users, then every user can only get as many PRBs as the worst user (using the highest transmit power). We can write
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ15_HTML.gif
(15)
An important observation is that this solution is also throughput fair in the case of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq110_HTML.gif (with the exception that power limited users would have smaller throughput). Otherwise ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq111_HTML.gif ) close users get higher throughput since the received power is higher and the interference is the same for all users in a cell.

3.4. Modified Resource Fair

The previous scheduler has the disadvantage that it may leave a lot of resources unused although close users would still be able to extend their bandwidth. Unfortunately, users at the cell edge with high propagation loss cannot make use of the spare bandwidth due to power limitation.
In another extreme solution we could try to always give every user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq112_HTML.gif its maximum allowed bandwidth https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq113_HTML.gif . If this does not exceed the available resources, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq114_HTML.gif , this is a viable approach. However, this will be relatively unlikely in reality since already a single close user could have enough transmit power to occupy more than https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq115_HTML.gif PRBs.
In this case we need to scale down the number of PRBs. The simplest solution would scale down all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq116_HTML.gif 's in the same way. However this would leave too much unfairness in the system. Instead we prefer scaling down large https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq117_HTML.gif 's and bring this new solution as close as possible to the resource fair case. We will call this solution modified resource fair although it is in general not resource fair. However, in annex A we will observe that this solution achieves the same fairness as the typical resource fair definition in the downlink.
We propose a simple iterative method which starts with the previous resource fair case. We define the indices https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq118_HTML.gif such that they address all users https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq119_HTML.gif in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq120_HTML.gif in ascending order with respect to https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq121_HTML.gif 's, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq122_HTML.gif addresses the worst user in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq123_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq124_HTML.gif addresses the second worst user, and so forth. We will formulate our algorithm as follows:
(1)
Initialize: https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq125_HTML.gif ; https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq126_HTML.gif
 
(2)
Abbreviate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq127_HTML.gif
 
(3)
if https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq128_HTML.gif
(a)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq129_HTML.gif
 
(b)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq130_HTML.gif else
 
(c)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq131_HTML.gif for all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq132_HTML.gif
 
(d)
exit
 
 
(4)
Increment https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq133_HTML.gif and go to step 2
 
In every iteration, we check whether the remaining resource budget https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq134_HTML.gif equally shared among the remaining https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq135_HTML.gif exceeds the PRB limit https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq136_HTML.gif of the worst of the remaining users https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq137_HTML.gif . If yes, the worst remaining user gets its maximum number of PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq138_HTML.gif , and we assign the remaining budget in the next iteration. Otherwise the remaining budget is equally shared among the remaining users, and we exit the algorithm.
Note again that in this solution the worst user gets the least amount of resources, but the maximum it can afford. With a high number of users this case will converge against the previous "Resource Fair" case.

3.5. Throughput Fair

In this section we try to approximate a throughput fair solution. We have already mentioned that the number of PRBs is limited for the users. Since the interference is the same for all users the throughput achievable by all users is determined by the worst user (in particular for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq139_HTML.gif ). The true throughput fair solution employs the rate function and writes as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ16_HTML.gif
(16)
for two users https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq140_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq141_HTML.gif in the same cell. Note that throughput fairness is required per cell. Unfortunately the SINRs are not known so far; recall that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq142_HTML.gif 's are needed to calculated scheduling probabilities and thereby the SINRs. Therefore we will give two different approximations in the following.
As a first approximation, we will do the simplifying assumption that the throughput is proportional to the SINR, that is, we assume linear rate function. From (7) we observe that the average SINR of a user within a certain cell is proportional to the received power (since the interference is cell specific). In this case the throughput fair criterion of the previous equation degenerates to
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ17_HTML.gif
(17)
Another approximation which is derived from Shannon's equation is
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ18_HTML.gif
(18)
The comparison of the two approximations is shown in Figure 2 where we have used https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq143_HTML.gif . The true relation obviously depends on the SINR range of the reference user (cf. legend). The linear approximation fits for very small SINR ranges; the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq144_HTML.gif approximation fits better for medium SINR ranges.
Both approximations have the very nice property that they only depend on the positions of the users within a cell and not on intercell interference or other cells in general. With those assumptions, we can formulate the throughput fair (approximated) solution in three steps.
First we assume that the worst user gets the maximum number of PRBs:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ19_HTML.gif
(19)
Next we derive the number of PRBs for all the other users in the cell by applying equation (17)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ20_HTML.gif
(20)
or (18)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ21_HTML.gif
(21)
Finally we need to check whether we have exceeded the resource limit. In this case, we have to scale down all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq145_HTML.gif 's by the same factor in order to fit into the available resources whilst maintaining the throughput fairness:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ22_HTML.gif
(22)

3.6. Quality of Service

A drawback of the previous methods is that we cannot define a target QoS or a user satisfaction level. Inherently the methods were based on the best effort and full buffer assumption. The users always have data to transmit on one hand; on the other hand they do not have to meet a certain target, that is, they are satisfied with whatever resources https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq146_HTML.gif they get.
For a variety of services a certain QoS target has to be met. For instance, users are only satisfied if they get a certain bit rate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq147_HTML.gif . If they get less, they are unsatisfied. On the other hand, they typically cannot transmit more than https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq148_HTML.gif , so the system will assign only the resources https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq149_HTML.gif such that the target rate is fulfilled, not more. Such a behavior is called constant bit rate(CBR) service.
Initially, let us assume that the SINRs are already known. We will resolve this assumption in the subsequent section. The approach is very similar to the approach in [13]. In order to achieve the target rate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq150_HTML.gif whilst observing the power (and therefore resource) limitation in uplink, we write the required average number of PRBs for user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq151_HTML.gif as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ23_HTML.gif
(23)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq152_HTML.gif is the rate function introduced in Section 2.5. It is important to observe that a user cannot be satisfied if the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq153_HTML.gif operator expires, irrespective of the traffic situation in the own cell (even if the user were alone). The only way to improve those users is to decrease the intercell interference, which requires modifications in the neighboring cell such as decreasing the P0 [21]. Note that any of those modifications is likely to reduce the QoS level in the neighboring cell.
A cell can be defined in overload if the sum of the required resources exceeds the available resources, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq154_HTML.gif . In this case contention control would drop some users (or, equivalently, admission control would not even have admitted some users). We assume that those control mechanisms work arbitrarily, that is, they do not prefer some (e.g., close) users and discriminate others (e.g., far users). This case can be modeled by applying the same scaling procedure as in (22):
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ24_HTML.gif
(24)
This scaling procedure would basically make every user unsatisfied. However note that the scheduling probabilities here are needed to calculate SINRs. Performance metrics will be discussed in Section 4. Alternatively, we could make use of admission control functionality here, which basically would select a subset https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq155_HTML.gif (and drops the other users) such that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq156_HTML.gif is fulfilled.
We would like to emphasize again that we have assumed that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq157_HTML.gif 's are already known. However, we actually need the scheduling probabilities to calculate the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq158_HTML.gif 's based on (7). So in contrast to the strict resource fair, modified resource fair and (approximated) throughput fair solutions of the previous sections, we unfortunately have not found a closed form solution for the QoS case. This problem is very similar to the downlink problem as described in [13].

3.7. Comparison with Real-World Schedulers

In the following we will discuss how real schedulers would map to the previously introduced strategies. The most popular scheduler is a proportional fair(PF) scheduler. The pure PF strategy is resource fair [18, 19]. However, unfortunately the PF definition in the uplink is not as straightforward as it is in the downlink due to power control and power limitation. Most of the uplink PF strategies in LTE will use adaptive transmission bandwidth and will be very close to the modified resource fair definition introduced in Section 3.4, when assuming full buffer/best effort traffic models (i.e., no further QoS constraints), compare, for example, [20]. Note that the scheduling gain, that is, the fact that the SINR conditioned on a user being scheduled gets better, goes into the throughput mapping discussed in Section 2.5 and not into the scheduling probabilities. Hence, PF and round robin strategies are equivalent from the perspective of scheduling probabilities (both are resource fair).
Furthermore, the PF strategies typically have to be extended with QoS constraints such as a target bit rate, minimum bit rate, or delay constraints. Those extended PF versions will come closer to the QoS scheduler described in Section 3.6. Once again, the reduced scheduling gain (through more QoS constraints) is considered in the throughput mapping, rather than in the scheduling probabilities.

3.8. Initialization of the SINRs

In this section we will propose 2 different solutions. Let us first recall the SINR definition from (7)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ25_HTML.gif
(25)
The first observation is that the abbreviated expectation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq159_HTML.gif is only cell specific and not user specific. Hence, for a first guess of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq160_HTML.gif 's according to (23) and (24), we only need to approximate a single value rather than https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq161_HTML.gif user-specific https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq162_HTML.gif 's, which seems to be a much simpler problem. If we are applying the framework in this paper to a dynamic simulator with a continuous time axis, we can simply take the guess of the expectation from the previous time step. Similarly, we can read that once we know the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq163_HTML.gif of one user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq164_HTML.gif (e.g., the worst user), we know all the others by the simple relation
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ26_HTML.gif
(26)
The advantage is that it might be easier to make a guess on the SINR since it is a relative number rather than a guess on the expectation which is an absolute number. In particular the SINR of the worst user in a cell is rather likely to be very small. So the second proposal is to set the SINR of the worst user in every cell to a predefined value https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq165_HTML.gif (e.g., 0 dB), and the other user's SINR in the same cell are derived from that according to (26). This method has the advantage that it also works with so-called snapshot- like simulators which do not have a time axis. In a dynamic simulator, this approach is probably less accurate than the first one.

4. Performance Metrics

So far, we have an (almost) analytical expression https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq166_HTML.gif for the average SINR of every user in an LTE uplink network. Furthermore, we have already discussed the average number https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq167_HTML.gif of assigned PRBs for different scheduling strategies. Note that in the QoS case the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq168_HTML.gif 's actually depend on the SINRs which are not known when calculating the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq169_HTML.gif 's. Hence, before calculating performance metrics we should update the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq170_HTML.gif 's with the more accurate values of the SINRs.
From these https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq171_HTML.gif 's and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq172_HTML.gif 's we now can start deriving several capacity metrics such as average cell throughput, throughput percentiles, or number of (un)satisfied users.

4.1. Throughput Metrics

In the simplest case, we calculate the user throughputs as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ27_HTML.gif
(27)
From those rates we can calculate a total network throughput, throughputs per cell, or throughput percentiles. In principle we could also check whether users are satisfied by comparing their data rates with the rate requirements https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq173_HTML.gif 's. However recall that in (24) we have scaled down the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq174_HTML.gif 's of all users in case of an overload. In this case, all users would fall below their https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq175_HTML.gif 's although in reality it might be sufficient to drop very few users to make the rest satisfied again. Furthermore, it would be interesting to have a quantitative notion of how much overloaded a cell is and how many users are unsatisfied in fact. So for the QoS case, we will define more appropriate performance metric in the following.

4.2. Overload and Unsatisfied Users

Exactly as in [13] we return to the required number of PRBs from (23) and define a virtual cell load
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ28_HTML.gif
(28)
which can exceed 1 thereby indicating the degree of overload. For instance, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq176_HTML.gif means a 10% overloaded cell, and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq177_HTML.gif means that the cell is double overloaded, that is, half of the users will be unsatisfied. Again assuming that an admission/contention control would exclude arbitrary users (not preferably cell edge users), we can write the number of unsatisfied users in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq178_HTML.gif as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ29_HTML.gif
(29)
This number accounts for dissatisfaction through overload. In addition, we will also have unsatisfied users through power limitation as already discussed in the context of (23), even if the virtual load is very small. We simply count their number in cell
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ30_HTML.gif
(30)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq179_HTML.gif returns the size of the set https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq180_HTML.gif . A further limitation on cell level is given by the fact that the number of users which can be scheduled at the same time is constrained by the available resources for control channels (physical downlink control channel PDCCH in LTE). Note that this can be a painful restriction in particular in the uplink, where the individual UE power budgets limit the ability of following an aggressive TDMA strategy. With our mathematical framework we can easily capture this limitation as well. Assume that the maximum number of schedulable users in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq181_HTML.gif per TTI is given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq182_HTML.gif . (This is a simplification. In LTE this is not a hard limit, but it depends on the user positions.) The control channel consumption is minimized by a scheduling strategy which would always assign the maximum number of resources https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq183_HTML.gif according to (13) to a scheduled user. This maximized the number of TTIs in which a user is not scheduled, that is, where it does not require any control resources. Hence, the (averaged) minimum number of required control channels required by user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq184_HTML.gif per TTI is
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ31_HTML.gif
(31)
using the required number of PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq185_HTML.gif from (23). Note that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq186_HTML.gif . Obviously, the control channels will definitely (even without any delay requirement) cause dissatisfaction in case
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ32_HTML.gif
(32)
Equivalent to the load dissatisfaction we will again assume that admission/contention control would exclude arbitrary users and thus we can define the number of unsatisfied users due to control channel limitation as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ33_HTML.gif
(33)
Finally we have to combine the three metrics https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq187_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq188_HTML.gif , and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq189_HTML.gif to a single number of unsatisfied users per cell. With our high level of abstraction this is quite challenging since the sets of load-, power-, and control-unsatisfied users might be overlapping. A heuristic approach would exclude users one by one (power-limited users first) and recalculate the metrics until dissatisfaction has disappeared. Another approach exploits the intuitive fact that the set of load- and control-limited users (i.e., the cell level metrics) are obviously fully overlapping. The set of power-limited users (user-level metric) will be rather disjoint. With those assumptions we approximate the total number of unsatisfied users in cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq190_HTML.gif as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ34_HTML.gif
(34)

5. Results

A dynamic system level simulator has been implemented based on the derivations in the previous chapters. In this section we will present some results with standard assumptions (such as full buffer traffic, proportional fair scheduler), and we will show that those are very close to other simulation results which have been agreed for by several companies in [9, 22]. Furthermore, we will present results with CBR traffic, and we will also look at an irregular network with SON adaptation of the power control parameters. Finally we will elaborate on the huge runtime performance.

5.1. Simulation Assumptions

We will use standard assumptions as proposed in [16], comprising a network of 19 LTE base stations with an intersite distance of 500 m, serving 57 hexagonal cells (sectors). Pathloss law, shadowing model, and horizontal beam pattern are also taken from [16], a vertical pattern is not used. The users are moving with a speed of 3 km/h, and they are handover to another cell if the received signal strength (measured on downlink reference signals) with respect to the new cell is 3 dB better than that with respect to the serving cell (handover hysteresis). One simulation step is 100 ms, that is, the network performance is evaluated 10 times a second.
We are using homogeneous P0 values of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq191_HTML.gif or https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq192_HTML.gif and a homogeneous https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq193_HTML.gif value of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq194_HTML.gif . The resulting distribution of transmit power per PRB is shown in Figure 3. Note that this distribution does not depend on the scheduling mechanism or traffic model since we record one power value for every user per simulation step.
It is obvious that the larger P0 setting of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq195_HTML.gif 52 dBm leads to higher transmit powers. In this case we can also identify the maximum transmit power of 23 dBm.

5.2. Full Buffer Traffic

We will start with the simple assumption of a full buffer traffic model and a modified resource fair scheduler as presented in Section 3.4. Users are uniformly dropped into the network area such that every cell serves an average of 10 users. The distribution of the user throughputs according to (27) is given in Figure 4.
As expected we observe slightly higher user throughputs with the larger P0 value. However, the difference between the curves is smaller in the lower part of the plot, since the power limitation is more critical with the smaller P0 value. The 5% percentiles (which is typically referred to as cell edge throughput) are420  kbps and503  kbps whereas the average cell throughputs are 7.3 Mbps and 8.5 Mbps, respectively. This is in very good agreement with the simulations in [9, 22]. The results of different companies are compared in [22] for the reference case which we have used as well. The cell throughput results are in the range between 6.3 Mbps and 1.01 Mbps, with an average of 8.6 Mbps (which is also the result of [9]). The cell edge results span from 100 kbps to 460 kbps with an average of 260 kbps. Obviously our results are a bit too optimistic in terms of cell edge throughput which could be a consequence of the neglected fast fading, and, even more important, of handover gain, which is included in our simulations with full mobility.

5.3. Constant Bit Rate Traffic

Next we will assume a constant bit rate traffic model and a QoS scheduler as presented in Section 3.6. Different target data rates are assumed, namely, 96 kbps, 256 kbps, and 512 kbps. Again, users are uniformly dropped into the network; however, the average number of users per cell is varied from 5 to 80. Let us first look at the percentage of unsatisfied users due to power limitation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq196_HTML.gif as given by expression (30) in Figure 5.
We observe the following behavior.
(i)
All curves reach a maximum and then do not grow any further. The reason is that the actual load is limited and cannot exceed 100%. So the interference will also not grow with the number of users, and the SINRs will not decrease.
 
(ii)
The (power) dissatisfaction level is larger for higher data rates. This is quite obvious.
 
(iii)
The (power) dissatisfaction level is larger for the larger https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq197_HTML.gif . With smaller P0, the users can afford more PRBs, compare (14), whereas the interference level goes down as well (note that the other cells will reduce P0 as well in our model). So the SINRs remain the same as long as we do not enter noise limited regimes.
 
(iv)
With 512 kbps and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq198_HTML.gif we even have a "dissatisfaction floor," that is, there will be power limited users even in an empty system. That is, high uplink data rates can only be supported with small https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq199_HTML.gif values (or by relaxing the ATB power constraint (14)).
 
Note that the previous figure did not take into account users which cannot be served due to the lack of bandwidth. Figure 6 shows the total number of unsatisfied users according to (34), that is, the sum of power- and load-unsatisfied users. Control limitation is not considered, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq200_HTML.gif .
Certainly we can recognize the aforementioned dissatisfaction floor for 512 kbps and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq201_HTML.gif in this figure. Otherwise, the impact of the P0 value is almost negligible since adding users beyond 100% virtual load obviously means load-unsatisfied users hiding the aforementioned limit for the dissatisfaction level due to the power constraint. If we target a typical overall dissatisfaction level of 5%, the uplink can satisfy 10, 21, and 56 user with 512 kbps, 256 kbps, and 96 kbps, respectively. The cell throughput with the smaller rates is around 5.4 Mbps whereas the 512 kbps case is slightly worse with 5.4 Mbps due to the more critical power limitation.
As expected the CBR capacity is significantly below the best effort capacity. However, the difference is smaller than in the downlink, since the power control compensates for a part of the SINR loss of cell edge users.

5.4. Heterogeneous Scenario

Next we will leave the homogeneous standard scenario and continue with a heterogeneous scenario with different cell sizes and nonuniform user concentrations. Figure 7 illustrated the scenario which has been proposed in [23]. The eNBs are located on an irregular grid, 8 users are dropped into every cell, and additional 42 users (i.e., 50 users in total) are dropped into cell no. 11 simulating a hot spot. All users use a CBR of 64 kbps. For every cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq202_HTML.gif an individual https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq203_HTML.gif is chosen such that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq204_HTML.gif operator in the power control equation (2) expires in roughly 5% of the cell area.
We will also look at load adaptive power control(LAPC) as proposed in [24] where the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq205_HTML.gif s are reduced in cells which only carry a small load. In the CBR model reducing https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq206_HTML.gif blows up the resource consumption since the resulting SINR loss has to be compensated by bandwidth. We use a very similar approach to [24] and update the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq207_HTML.gif at time step https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq208_HTML.gif depending on the previous value https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq209_HTML.gif and the previous virtual load https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq210_HTML.gif (note that this equation is in dB scale):
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ35_HTML.gif
(35)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq211_HTML.gif is the virtual load which we are targeting. In theory we may want to target 100%; however, experience has shown that a margin should be left for handover users so that we will use https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq212_HTML.gif . The rule means that we increase the current https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq213_HTML.gif if the load is above target, and we decrease it if the load is below the target; however, we will not increase the initial https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq214_HTML.gif which has been defined above. Note that this automatic adaptation of a cell parameter can already be considered as a SON mechanism.
Figure 8 depicts the virtual loads in the overloaded cell no.11 and its neighbors over time where we have switched on the LAPC at https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq215_HTML.gif . Before that, the virtual loads are rather small (except the overloaded cell no.11) and different in every cell depending on the exact position of the users and the cell shape/size. After switching on the LAPC the virtual load in all low-loaded cells approaches the target https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq216_HTML.gif .
The time characteristics of the corresponding https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq217_HTML.gif s are shown in Figure 9. Without LAPC we can observe that the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq218_HTML.gif s depend on the cell size. Large cells have small https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq219_HTML.gif s and vice versa (due to the aforementioned 5% rule). After switching on LAPC, the low-loaded cells reduce their https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq220_HTML.gif s whereas cell no.11 does not change it.
Now let us look at the impact of the LAPC on the distribution of the interference over thermal(IoT) values. Those are based on the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq221_HTML.gif samples used for the Monte Carlo approach defined in (10); the exact definition of the (instantaneous) IoT is given by
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ36_HTML.gif
(36)
leading to https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq222_HTML.gif IoT samples per cell. Figure 10 shows the cumulative distribution function of the IoT values (in dB) in cell no.11 before and after switching on LAPC, and it also displays the linear and the harmonic averages of the IoT given by
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ37_HTML.gif
(37)
With LAPC the CDF is steeper since the users spread their data rate over a larger bandwidth leaving less PRBs unused (without interference) and leading to a smoother interference. The interference of an individual user per PRB certainly goes down significantly (with the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq223_HTML.gif reduction); however, it has to occupy more PRBs to reach its CBR. Still, the linear average of the IoT is smaller with LAPC since the rate function is concave over a wide area meaning that decreasing the power can be compensated by a smaller increase of the bandwidth. However, the harmonic average shows a surprising picture. The harmonic average decreases with LAPC which is a consequence of the larger variance of the IoT. Note that we have clearly shown in Section 2.3 that the harmonic average is actually the relevant measure. This also manifests in the distribution of the average user SINRs defined in (7) which are shown in Figure 11 for the overloaded cells.
The LAPC has degraded the SINR in the overloaded cell even though the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq224_HTML.gif has not been reduced there since the more fluctuating interference obviously offers more potential for the link adaptation (which is assumed to be ideal in our model). This is also visible in the virtual load of the overloaded cell no.11 in Figure 8 which has slightly been increased by LAPC (as a result of the decreased SINRs). This is a contrast to the results in [24] where LAPC helps to improve the system. Fluctuating interference has a negative impact on link adaptation (i.e., selection of modulation and coding schemes, scheduling, etc.). In other words, smoothening the interference through LAPC will improve link adaptation. Unfortunately this effect is not covered in our simplified model, where link adaptation is always assumed to be ideal. Hence, our model can exploit the aforementioned potential offered by fluctuating interference, which is not the case in reality.
Although we have gained important insights by this analysis, it also reveals a current limitation of the model. A remedy could be based on the principles of [25], where the rate function is elaborated by the introduction of a correlation between the SINR at the moment of choosing the modulation and coding scheme and the moment of applying it (where the interference might have changed). This correlation would increase through LAPC.

5.5. Simulation Runtime

Finally we will look at the runtime performance of the simulation. Figure 12 shows the simulation time for one simulation step versus the average number of users per cell. It turned out that the number of samples used for the Monte Carlo integration in (10) has significant impact. As already mentioned, convergence is achieved for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq225_HTML.gif . Fortunately we observe that further reduction of this number does not bring additional runtime benefit. The increase is linear with respect to the number of users, which is not surprising. With 50 users per cell and a simulation step of 1 sec, which is sufficient for many applications, we are a factor 2 above real-time. Recall that we have a fully heterogeneous 57-cell network, and no homogeneous properties are exploited.

6. Conclusions

We have presented a very efficient modeling approach for uplink investigations focussing on the LTE standard. QoS and radio resource management (which typically work on a millisecond time scale) are modeled in a very abstract but still accurate way such that the essential behavior is still included. By those means we can decrease simulation runtime far below real-time.
This is in particular helpful, even necessary, for simulations covering long time intervals. The most important applications are investigations on self-optimizing networks since SON mechanisms are typically very slow control loops and converge over hours or even days. Conventional system level simulators cannot serve this purpose. On the other hand, QoS and mobility issues are of utmost importance and can not be neglected when studying SON which makes static tools inappropriate as well. We are closing the gap between too slow but elaborate system level simulators with full mobility and QoS support on one side and rough static simulators which can lead to very fast results.
We have seen that uplink modeling is much more complicated than downlink modeling. The key differences are the uplink power control (including the associated individual UE power budgets) and the multiple-access structure of the uplink (leading to extremely fluctuating interference). We have derived an average uplink SINR which is equivalent to the downlink SINR which is typically intuitively used. We have observed that the uplink interference has to be averaged in a harmonic way.
Different traffic/scheduler assumptions have been discussed. Again in contrast to downlink, there is no unique definition of a resource fair scheduler in the full buffer case. We have given two solutions called strict and modified resource fair. Furthermore, throughput fair solutions as well as CBR solutions targeting a given bit rate have been defined. In order to evaluate the system performance we have discussed uplink satisfaction in the CBR case. In addition to load limitation, we have observed that satisfaction due to power limitation and due to control channel limitation is highly relevant in uplink, too.
Limitations of the abstract models have been addressed as well. In particular, RRM details such as the exact algorithms for MCS selection, multi-user diversity gains, and imperfect channel estimation are hidden behind the abstract models (although the essence of fairness issues is considered). Furthermore, the capability to consider more elaborate traffic models (different from pure CBR and pure best effort) is limited. We are giving an outlook on possible extensions in Appendix B.
We have given simulation results using the derived modeling approach. In the specified LTE test cases our results match very well the typical performance assumptions. We have also given results for the CBR case using different target bit rates and analyzed the impact of power limitation. Finally we were looking at a heterogeneous scenario with different cell sizes and non-uniform user placement. We have considered load-adaptive power control as an example for a SON mechanism. This scenario has revealed some limitations of our modeling approach which have to be improved in the future.
In this paper we have only looked at a small subset of the proposed SON use cases since the focus was on the introduction of the framework. Certainly the model can be used for all other SON use cases as well, such as load balancing, coverage and capacity optimization, or mobility robustness optimization.

Appendices

In this section we will compare the uplink fairness with the downlink fairness. We will show that themodified resource fair scheduler in the uplink achieves a comparable fairness as a typical resource fair scheduler in downlink.
In downlink, the SINRs on a PRB degrade towards the cell edge more severe than the pathloss law, since the interference grows in addition. With a resource fair scheduler, the throughputs behave accordingly. However, we can improve cell edge users arbitrarily by assigning them more PRBs (nonresource fair scheduling).
In the uplink, the degradation towards the cell edge is different. By purely looking at the power control equation (2), bandwidth limitation (13), and SINR definition (7), we make the following observations.
(1)
The uplink interference is induced at the eNB antennas and therefore is the same for all users (as long as we do not assume intercell interference coordination).
 
(2)
Assuming no power control https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq226_HTML.gif (resulting in strict resource fairness, that is, one PRB per user), that is, each UE transmits with maximum power, the SINRs per PRB would degrade with the pathloss. Note that this is still "fairer" as in the downlink (where interference increases in addition).
 
(3)
Assuming power control with full pathloss compensation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq227_HTML.gif (and adaptive transmission bandwidth), that is, all UEs are received with the same power per PRB, the SINRs per PRB would be the same for all UEs, but the assigned bandwidth degrades with the pathloss, copmared with (14) unless the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq228_HTML.gif s are rather small. So the throughputs degrade with the pathloss (which is a little steeper than the upper case due to concavity of the rate function).
 
(4)
Assuming power control with fractional pathloss compensation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq229_HTML.gif (and adaptive transmission bandwidth), the SINRs per PRB would degrade less severely than the pathloss; however, the assigned bandwidth degrades with the "rest" of the pathloss law. So in total the throughputs again will degrade with pathloss, with the slope being in between the upper two cases.
 
So in either case the throughputs will degrade with the pathloss, either via SINR degradation (for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq230_HTML.gif ) and / or via the ATB scheduler. The degradation will be similar to the downlink. All cases refer to the definition of the modified resource fair scheduler.The strict resource fair solution will lead to more fairness, that is, less throughput degradation for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq231_HTML.gif . For the special case of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq232_HTML.gif strict resource fairness even leads to throughput fairness.
Despite similar fairness behavior, the uplink scheduler has much less degrees of freedom to trade throughput among the users (due to individual power budgets). The only way to extend the strict throughput limit of cell edge users is to reduce the intercell interference which unfortunately lies outside the responsibility of the serving cell. Even more severe, the neighbors can only decrease interference by degrading their own users. Hence, whereas in downlink we can trade throughputs between users, in uplink we need to trade throughputs between cells.

B. Traffic Assumptions

We have already observed that the individual power limitation in the uplink results in a bandwidth limitation https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq233_HTML.gif (the number of PRBs for edge users is limited) and thereby in a tight data rate limit. This is different from the downlink, where basically each PRB comes along with its own power budget, so that assigning more PRBs automatically means assigning more power.
As a consequence, in the uplink we can only guarantee small data rates to cell edge users. So if we want to assume a common CBR model for all users, it has to be very small such that the edge users have a fair chance to achieve it. With such a setting there are basically 3 different methods how capacity limit could be achieved.
( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq234_HTML.gif ) Common CBR
The straightforward solution is to consider a large number of users. This would increase simulation runtime, and large data rates would not occur at all.
( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq235_HTML.gif ) Common CBR Extended with Full Buffer
With a smaller number of users we could think about distributing the excess capacity (i.e., PRBs) among those users who still can afford more PRBs. Basically this would be a mixture of CBR and full buffer model, which can be referred to as guaranteed bit rate(GBR) model. The drawback is that we would need another metric (in addition to the satisfied users due to load and power) accounting for users with higher throughput (i.e., 95% percentile). Furthermore we have to set a rule how the excess capacity is distributed among the users.
( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq236_HTML.gif ) User-Specific CBR
Finally, we could set different bit rate requirements https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq237_HTML.gif for the users, for example, depending on the pathloss. Edge users would get small https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq238_HTML.gif 's, and close users would get higher https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq239_HTML.gif 's. This might be the most convenient solution and probably the most realistic one as well. However, we need to find an appropriate rule to set the individual rate requirements https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq240_HTML.gif 's, which is absolutely not straightforward. We will propose a simple mechanism with the following characteristics.
(a)
We define a minimum data rate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq241_HTML.gif which lower-bounds all rate requirements, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq242_HTML.gif .
 
(b)
The rate requirement for the worst user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq243_HTML.gif (i.e., with largest pathloss) in every cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq244_HTML.gif is set to the minimum rate, that is,
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ38_HTML.gif
(B1)
 
(c)
The rate requirements for the other users are upscaled according to the pathloss relation to the worst user https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq245_HTML.gif
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ39_HTML.gif
(B2)
The fairness parameter https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq246_HTML.gif can be considered as "slope" of the rate requirements. https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq247_HTML.gif obviously means no slope, that is, the same rate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq248_HTML.gif for all users which converges to the common CBR solution. There is an interesting relation to the power control parameter https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq249_HTML.gif ; setting https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq250_HTML.gif will lead approximately to a resource fair behavior, since the rate increase would roughly be compensated by the increase in received power through fractional power control; this would be exactly true for a linear relation between SINR and throughput and if we neglect the power limitation. https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq251_HTML.gif is the most aggressive setting which makes it approximately equally tough for every user to achieve its target.
 
For this special case of user-specific CBR setting, we will now propose a special initialization for the SINRs. Recall that the SINRs depend on the average number of PRBs https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq252_HTML.gif (or scheduling probabilities, resp.); however, the performance metrics depend on the SINRs in turn. For 3 artificial cases we have proposed simple ways to calculate the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq253_HTML.gif 's without needing the SINRs in Sections 3.3, 3.4, and 3.5. The proposal is a slight modification of the idea for the TP-fair case in Section 3.5. In every cell https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq254_HTML.gif we start with an initial guess https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq255_HTML.gif for the worst user's https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq256_HTML.gif . From this guess we approximate the resource consumption of
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_Equ40_HTML.gif
(B3)
if we neglect power limited users. As discussed before we observe
(i)
the resource fair behavior https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq257_HTML.gif for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq258_HTML.gif ,
 
(ii)
smaller resource consumption https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq259_HTML.gif for close users towards the throughput fair solution https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq260_HTML.gif ,
 
(iii)
larger resource consumption https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq261_HTML.gif for close users towards the aggressive solutions https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq262_HTML.gif ;with this setting, we can initialize the SINR calculation for the user-specific CBR case depending only on a single parameter https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq263_HTML.gif .
 
(4) Summary
We will now summarize the required hooks for the traffic configuration with user-specific CBRs.
(i)
The first is a minimum rate requirement https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq264_HTML.gif . We propose values between 64kbps and 96kbps. Higher values will already become critical for cell edge users.
 
(ii)
The second is a slope https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq265_HTML.gif for the rate requirements. https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq266_HTML.gif means throughput fair behavior; smaller values increase the rate requirements for closer users depending on their pathloss relation to the worst user, thereby forcing more load in the system. We propose a setting between 0 and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq267_HTML.gif .
 
(iii)
The third is a guess https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq268_HTML.gif for SINR initialization. A first proposal is https://static-content.springer.com/image/art%3A10.1155%2F2010%2F282465/MediaObjects/13638_2010_Article_1851_IEq269_HTML.gif ; however, this requires further study.
 
Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License ( https://​creativecommons.​org/​licenses/​by/​2.​0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Literatur
2.
Zurück zum Zitat Döttling M, Viering I: Challenges in mobile network operation: towards self-optimizing networks. Proceedings of IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP '09), April 2009, Taipei, Taiwan 3609-3612. Döttling M, Viering I: Challenges in mobile network operation: towards self-optimizing networks. Proceedings of IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP '09), April 2009, Taipei, Taiwan 3609-3612.
3.
Zurück zum Zitat Amirijoo M, Litjens R, Spaey K, Döttling M, Jansen T, Scully N, Türke U: Use cases, requirements and assessment criteria for future self-organising radio access networks. Proceedings of the 3rd International Workshop on Self-Organizing Systems (IWSOS '08), December 2008 275-280. Amirijoo M, Litjens R, Spaey K, Döttling M, Jansen T, Scully N, Türke U: Use cases, requirements and assessment criteria for future self-organising radio access networks. Proceedings of the 3rd International Workshop on Self-Organizing Systems (IWSOS '08), December 2008 275-280.
6.
Zurück zum Zitat Türke U, Koonert M: Advanced site configuration techniques for automatic UMTS radio network design. Proceedings of the 61st IEEE Vehicular Technology Conference, May-June 2005, Stockholm, Sweden 3: 1960-1964. Türke U, Koonert M: Advanced site configuration techniques for automatic UMTS radio network design. Proceedings of the 61st IEEE Vehicular Technology Conference, May-June 2005, Stockholm, Sweden 3: 1960-1964.
7.
Zurück zum Zitat Lobinger A, Stefanski S, Jansen T, Balan I: Load balancing in downlink LTE self-optimizing networks. Proceedings of the 71st IEEE Vehicular Technology Conference, May 2010, Taipei, Taiwan Lobinger A, Stefanski S, Jansen T, Balan I: Load balancing in downlink LTE self-optimizing networks. Proceedings of the 71st IEEE Vehicular Technology Conference, May 2010, Taipei, Taiwan
8.
Zurück zum Zitat Brueninghaus K, Astélyt D, Salzer T, Visuri S, Alexiou A, Karger S, Seraji G-A: Link performance models for system level simulations of broadband radio access systems. Proceedings of the 16th IEEE International Symposium on Personal, Indoor and Mobile Radio Communications (PIMRC '05), September 2005, Berlin, Germany 4: 2306-2311. Brueninghaus K, Astélyt D, Salzer T, Visuri S, Alexiou A, Karger S, Seraji G-A: Link performance models for system level simulations of broadband radio access systems. Proceedings of the 16th IEEE International Symposium on Personal, Indoor and Mobile Radio Communications (PIMRC '05), September 2005, Berlin, Germany 4: 2306-2311.
9.
Zurück zum Zitat Rinne M, Kuusela M, Tuomaala E, Kinnunen P, Kovács I, Pajukoski K, Ojala J: A performance summary of the evolved 3G (E-UTRA) for voice over internet and best effort traffic. IEEE Transactions on Vehicular Technology 2009, 58(7):3661-3673.CrossRef Rinne M, Kuusela M, Tuomaala E, Kinnunen P, Kovács I, Pajukoski K, Ojala J: A performance summary of the evolved 3G (E-UTRA) for voice over internet and best effort traffic. IEEE Transactions on Vehicular Technology 2009, 58(7):3661-3673.CrossRef
10.
Zurück zum Zitat Simonsson A, Furuskär A: Uplink power control in LTE—overview and performance. Proceedings of the 68th IEEE Vehicular Technology Conference (VTC '08), September 2008, Calgary, Canada Simonsson A, Furuskär A: Uplink power control in LTE—overview and performance. Proceedings of the 68th IEEE Vehicular Technology Conference (VTC '08), September 2008, Calgary, Canada
11.
Zurück zum Zitat Majewski K, Koonert M: Conservative cell load approximation for radio networks with shannon channels and its application to LTE network planning. Proceedings of the 6th Advanced International Conference on Telecommunications, May 2010, Barcelona, Spain Majewski K, Koonert M: Conservative cell load approximation for radio networks with shannon channels and its application to LTE network planning. Proceedings of the 6th Advanced International Conference on Telecommunications, May 2010, Barcelona, Spain
12.
Zurück zum Zitat 3GPP : Simulation assumptions and parameters for FDD HeNB RF requirements. TDoc R4-092042, RAN WG4 Meeting #51, San Francisco, Calif, USA, May 2009, http://www.3gpp.org/ 3GPP : Simulation assumptions and parameters for FDD HeNB RF requirements. TDoc R4-092042, RAN WG4 Meeting #51, San Francisco, Calif, USA, May 2009, http://​www.​3gpp.​org/​
13.
Zurück zum Zitat Viering I, Döttling M, Lobinger A: A mathematical perspective of self-optimizing wireless networks. Proceedings of IEEE International Conference on Communications (ICC '09), June 2009, Dresden, Germany Viering I, Döttling M, Lobinger A: A mathematical perspective of self-optimizing wireless networks. Proceedings of IEEE International Conference on Communications (ICC '09), June 2009, Dresden, Germany
17.
Zurück zum Zitat Zerlin B, Ivrlač M, Utschick W, Nossek J, Viering I, Klein A: Joint optimization of radio parameters in HSDPA. Proceedings of the 61st IEEE Vehicular Technology Conference (VTC '05), May 2005, Stockholm, Sweden 1: 295-299. Zerlin B, Ivrlač M, Utschick W, Nossek J, Viering I, Klein A: Joint optimization of radio parameters in HSDPA. Proceedings of the 61st IEEE Vehicular Technology Conference (VTC '05), May 2005, Stockholm, Sweden 1: 295-299.
18.
Zurück zum Zitat Tse D, Viswanath P: Fundamentals of Wireless Communications. 1st edition. Cambridge University Press, Cambridge, UK; 2005.CrossRefMATH Tse D, Viswanath P: Fundamentals of Wireless Communications. 1st edition. Cambridge University Press, Cambridge, UK; 2005.CrossRefMATH
19.
Zurück zum Zitat Viswanath P, Tse DNC, Laroia R: Opportunistic beamforming using dumb antennas. IEEE Transactions on Information Theory 2002, 48(6):1277-1294. 10.1109/TIT.2002.1003822MathSciNetCrossRefMATH Viswanath P, Tse DNC, Laroia R: Opportunistic beamforming using dumb antennas. IEEE Transactions on Information Theory 2002, 48(6):1277-1294. 10.1109/TIT.2002.1003822MathSciNetCrossRefMATH
20.
Zurück zum Zitat Calabrese FD, Rosa C, Anas M, Michaelsen PH, Pedersen KI, Mogensen PE: Adaptive transmission bandwidth based packet scheduling for LTE uplink. Proceedings of the 68th IEEE Vehicular Technology Conference (VTC '08), September 2008, Calgary, Canada Calabrese FD, Rosa C, Anas M, Michaelsen PH, Pedersen KI, Mogensen PE: Adaptive transmission bandwidth based packet scheduling for LTE uplink. Proceedings of the 68th IEEE Vehicular Technology Conference (VTC '08), September 2008, Calgary, Canada
21.
Zurück zum Zitat Boussif M, Quintero N, Calabrese FD, Rosa C, Wigard J: Interference based power control performance In LTE Uplink. Proceedings of IEEE International Symposium on Wireless Communication Systems (ISWCS '08), October 2008, Reykjavik, Iceland 698-702. Boussif M, Quintero N, Calabrese FD, Rosa C, Wigard J: Interference based power control performance In LTE Uplink. Proceedings of IEEE International Symposium on Wireless Communication Systems (ISWCS '08), October 2008, Reykjavik, Iceland 698-702.
23.
Zurück zum Zitat Turkka J, Lobinger A: Non-regular layout for cellular network system simulations. Proceedings of the International Symposium on Personal, Indoor and Mobile Radio Communications (PIMRC '10), September 2010, Istanbul, Turkey Turkka J, Lobinger A: Non-regular layout for cellular network system simulations. Proceedings of the International Symposium on Personal, Indoor and Mobile Radio Communications (PIMRC '10), September 2010, Istanbul, Turkey
24.
Zurück zum Zitat Boussif M, Rosa C, Wigard J: Load adaptive power control in LTE uplink. European Wireless, April 2010, Lucca, Italy Boussif M, Rosa C, Wigard J: Load adaptive power control in LTE uplink. European Wireless, April 2010, Lucca, Italy
25.
Zurück zum Zitat Castañeda M, Ivrlač MT, Nossek JA, Viering I, Klein A: Outdated uplink adaptation due to changes in the scheduling decisions in interfering cells. Proceedings of IEEE International Conference on Communications (ICC '08), May 2008, Beijing, China 4948-4952. Castañeda M, Ivrlač MT, Nossek JA, Viering I, Klein A: Outdated uplink adaptation due to changes in the scheduling decisions in interfering cells. Proceedings of IEEE International Conference on Communications (ICC '08), May 2008, Beijing, China 4948-4952.
Metadaten
Titel
Efficient Uplink Modeling for Dynamic System-Level Simulations of Cellular and Mobile Networks
verfasst von
Ingo Viering
Andreas Lobinger
Szymon Stefanski
Publikationsdatum
01.12.2010
Verlag
Springer International Publishing
DOI
https://doi.org/10.1155/2010/282465

Weitere Artikel der Ausgabe 1/2010

EURASIP Journal on Wireless Communications and Networking 1/2010 Zur Ausgabe