integrated multiscale models for the optimal integration of …€¦ ·  · 2017-12-08integrated...

27
Integrated multiscale models for the optimal integration of renewable and distributed electricity generation Panos Parpas February, 2010 Abstract This paper identifies the challenges involved with addressing the integration of renewable and distributed electricity generation within the existing energy infrastructure. We outline a tentative framework to address these challenges using a multiscale modeling methodology. The integrating theme of the proposed methodology is suggested by the space–time separation grounded in the en- gineering and physical properties of large power networks. The issue of multiple scales is addressed using the framework of singular perturbation theory for the control of continuous time problems. We show that the uncertainties that exist in energy planning problems, in particular load demand uncertainty and uncertainties due to the power source, are highly structured. This structure can be exploited to a great advantage that provides insight into the nature of the problem. At the same time this structure can be used to perform a dimensionality reduction to the original model and therefore allowing useful computation to be performed. We discuss two different case studies. The first case study is concerned with the coupling of a capacity expansion model with a unit commitment model. The question addressed here is how to incorporate hourly dynamics in the capacity expansion model. The latter model is mainly concerned with investment decisions that are taken only every few years and their effects are long lived. The hourly dynamics describe electricity load demand and the intermittency of wind generation. This case study is thus concerned with the large scale integration of renewables in an existing portfolio of power generation. We show how to link the two models along their state dimension. The second case study addresses some of the challenges when considering the large scale integration of distributed generation or microgrids. This problem has a spatial aspect that is not present in the first case study and so we use it as a * Energy Initiative, M.I.T, Cambridge, MA. 02139. [email protected] 1

Upload: ngokien

Post on 20-May-2018

217 views

Category:

Documents


3 download

TRANSCRIPT

Integrated multiscale models for the optimal integration of renewable

and distributed electricity generation

Panos Parpas∗

February, 2010

Abstract

This paper identifies the challenges involved with addressing the integration of renewable and

distributed electricity generation within the existing energy infrastructure. We outline a tentative

framework to address these challenges using a multiscale modeling methodology. The integrating

theme of the proposed methodology is suggested by the space–time separation grounded in the en-

gineering and physical properties of large power networks. The issue of multiple scales is addressed

using the framework of singular perturbation theory for the control of continuous time problems.

We show that the uncertainties that exist in energy planning problems, in particular load demand

uncertainty and uncertainties due to the power source, are highly structured. This structure can

be exploited to a great advantage that provides insight into the nature of the problem. At the

same time this structure can be used to perform a dimensionality reduction to the original model

and therefore allowing useful computation to be performed. We discuss two different case studies.

The first case study is concerned with the coupling of a capacity expansion model with a unit

commitment model. The question addressed here is how to incorporate hourly dynamics in the

capacity expansion model. The latter model is mainly concerned with investment decisions that are

taken only every few years and their effects are long lived. The hourly dynamics describe electricity

load demand and the intermittency of wind generation. This case study is thus concerned with

the large scale integration of renewables in an existing portfolio of power generation. We show

how to link the two models along their state dimension. The second case study addresses some of

the challenges when considering the large scale integration of distributed generation or microgrids.

This problem has a spatial aspect that is not present in the first case study and so we use it as a

∗Energy Initiative, M.I.T, Cambridge, MA. 02139. [email protected]

1

template to illustrate how one could approach problems with a strong spatial component using a

multiscale framework.

1 Introduction

Due to the manifold of technoeconomic issues that arise in energy planning, the electricity industry

has for a long time relied on complex optimization models to perform its planning. Examples of such

models include unit commitment (decisions on which generators to use), economic dispatch (decide

the optimal level of use for each generator), capacity expansion (decide the optimal generation mix

to meet demand), hydro/thermal scheduling models, optimal power flow models and so on. While

there exists a large body of literature to address the range of challenges faced by the energy industry,

the environment in which these models are used is becoming increasingly complex. This complexity

is driven primarily by two factors. The first factor is the deregulation and re-regulation of electricity

markets. The second driver is the increasing pressure placed on the electricity industry to address

the problem of meeting demand in a sustainable manner. In order to deal with this complexity either

simplified models or heuristic methods are typically used. These approximations, are to a large extend,

uncontrolled and may have a substantial impact on the nature of the obtained solution. In particular

most models ignore the fact that many features, that have important ramifications for the selection of

the optimal policy, exist along multiple temporal and spatial scales. The impact of the optimization

methodology, and the potential inefficiencies of the old heuristics can be seen by two examples taken

from [Bloom, 2009]. The first is from Red Electrica de Espana (rEE), the Spanish national power

distributor. They switched from using heuristic models to an exact optimization methodology and

saved about 50k€-100k€ daily, and reduced emissions by about 2.5% annually (about 100,000 tons of

CO2). A second examples is a large system operator in the east of the United States that saved about

$200 million annually by solving the unit commitment problem by a branch and bound algorithm as

opposed to the heuristic method they used before. These examples illustrate the impact that ad hoc

heuristics can have on the energy industry. Incorporating large amounts of intermittent renewable

generation resources or the integration of distributed sources of electricity generation, makes an already

complex problem more complex. It is therefore important to have robust models that can scale, both

computationally and conceptually, with the increasingly complex issues faced in the energy industry.

2

This paper identifies the challenges involved with addressing the integration of renewable and

distributed electricity generation within the existing energy infrastructure. We outline a tentative

framework to address these challenges using a multiscale modeling methodology. The integrating

theme of the proposed methodology is suggested by the space–time separation grounded in the engi-

neering and physical properties of large power networks. Before we delve into the proposed modeling

methods it is necessary to decompose the implications of the two drivers and sources of the complexity

in which planning decision need to be taken. We refer the interested reader to [Claudio et al., 2008]

for a more complete discussion of these issues. Below we summarize the issues that have repercussions

on the planning models discussed in this paper.

2 Background

The deregulation of the industry means that utilities now have to compete in an open market envi-

ronment. The degree of deregulation varies from country to country, and from state to state in the

US. Despite these variations between regions, as well as the fact that real-time system operation is

still largely a centralized activity, deregulation has completely changed the way electricity generation

and long term planning are performed. In a regulated environment the utilities would pass on the

costs of any risks in generation costs directly to consumers. This approach is no longer possible in an

open market. Consequently, utilities see generation plants as an asset that needs to be risk managed

and optimally utilized. The focus is no longer on cost minimization, but on how to deliver a quality

product in an economically viable manner. However, electricity is different from other assets since

large scale storage is not economically feasible and so it needs to be generated and transmitted at the

same time it is needed. Thus the utility needs to maintain an equilibrium between generation and

demand in a way that ensures the safety and stability of the system, while maximizing asset utilization

at the same time. An important implication of the non-storability of electricity for long term planing,

is that it is not enough to know the total electricity needed, but also when it is needed. A system

under constant demand has dramatically different optimal design than a system facing a demand pro-

file with substantial variations over time. This property of electrical power system design remains the

same even if the total demand required is the same but the demand profile over time is different. One

way to address both the increasing volatility in generation costs as well as the way utilities respond

3

to variations in load is via investing in a portfolio of generation facilities. For example, nuclear plants

have large upfront investment costs, generally lower fuel costs, and are ideal for serving base load.

On the other end of the spectrum, gas turbines are cheap to build but have high fuel costs, and can

be used to meet peak demand. Utilities have also ventured, with mixed results [Parsons, 2008], into

trading. The reason behind the change of business model was either for hedging, risk management

purposes or so as to establish an alternative source of income. Thus considering a large and diverse

number of technologies and economic activities is an important facet of long term planning in the

energy sector. In a nutshell, the consequences of deregulation to capacity expansion decisions are the

increased need to incorporate robust methods for handling uncertainty and the importance of model-

ing load variations. Both of these factors are important if a substantial increase in system efficiency is

to be achieved. All these new facets of the capacity expansion problem need to be considered in view

of many possible current and evolving technologies. It is no longer possible to model each decision

in isolation. A system wide view encompassing not only physical, operational, and safety constraints

but also economic and, as it will be discussed next, environmental constraints as well, is required.

Until recently, and to a large extend even now, the environmental costs of producing electricity have

not been incorporated fully into the price consumers see. The process of internalizing these costs into

the price is currently under way. This requires utilities to think differently about some technologies,

and their role in helping make the transition into a system that can grow sustainably. For example, if

a utility wants to investigate the possibility of including a large amount of wind generation capacity

into their power portfolio, then due to the intermittent nature of wind, daily variations in load and

daily variations in wind speeds become more important. These daily variations will affect the way the

utility manages its reserves, and therefore effects the whole portfolio of generating capacity. Since the

output from a wind turbine increases by the cube of wind speed, using average wind speed will always

underestimate the amount of output available from wind and this can lead to an inefficient system

design. Again a system wide view needs to be taken. Currently, there exist little understanding on

how to, rigorously, capture these daily effects into long term planning.

The rest of this paper is structured as follows: in Section 3 we describe the general framework

we propose in order to capture some of the multiscale challenges described above. In Section 4 we

discuss a case study that couples capacity expansion with a unit commitment model. The results from

4

section 4 are taken from [Parpas and Webster, 2010]. In Section 5 we attempt to apply the methods

from [Parpas and Webster, 2010] to a distributed network planning problem. The analysis in 5 is both

speculative and exploratory in nature. We conclude in Section 6.

3 Overview of the proposed framework

In Figure 1 we illustrate the multiscale nature of decisions that need to be taken in the power systems

sector. In an intraday scale a utility will be concerned with the safe operation of the power generating

plants in order to meet demand on a second by second basis. Problems such as voltage control

and frequency control are used for this purpose. As we move up on the temporal scale, pricing,

risk management and capacity expansion problems are considered. The last box shows the range

of issues that are important on the larger socioeconomic scale that power systems belongs to. The

Info

rm

ati

on

Arr

iva

l

Seconds /

Weeks< Second

Weeks/

Years

Years /

Decades

Engineering /

Control

Unit commitment

Economic Dispatch

Pricing

Stability Control

Frequency Control

Voltage Control

Physics Engineering /

Economics

Generation Expansion

Network Expansion

Maintenance

Fuel Management

Power Management

Risk Management

Pricing

Bidding Strategies

Macroeconomics /

Development

Equilibrium models

System dynamics

Environment

Security

Policy

Size

Generator /

Transmission line

Plant, Region

Country

World / Economy

Figure 1: Multiscale models and decision on the power system sector.

deregulation and environmental concerns described in Section 2 make it necessary for utilities and

policy makers to use models that can bridge the scales between the different facets of power system

planning. Decisions taken by one agent in one scale (e.g. what technology mix to invest in) affect issues

that have their dynamics rooted at a different scale (e.g. environment). What an economist might call

an externality (e.g. greenhouse gas emissions) can also be thought of as a problem of interacting scales

[Cash and Moser, 2000]. This is because the environmental phenomenon (greenhouse gas emissions)

is managed at a scale (utility companies) that have no or little incentives to incorporate the temporal

and spatial dynamics of the phenomenon into their planning process. At the same time, policy

5

instruments (e.g. carbon taxes) are designed with the purpose of mitigating long term environmental

impacts but also have an impact on the choice of technologies and amount of innovation at the firm

level. This type of technological change is not captured in large economic models concerned with

the long time scales. The problem is more complicated than what is shown in Figure 1. Agents on

different scales use different models and with different resolutions. For example, an economist may

use a detailed model of the US economy but only an approximated model for the EU economy. This

multi-model environment introduces additional complications that is beyond the scope of this paper.

Hitherto the problem of bridging the scales has been handled using ad-hoc methods e.g. [Wing, 2006,

Bohringer and Rutherford, 2008]. The objective of this paper is to describe a methodology to address

some aspects of this problem.

The modeling approach proposed in this paper is based on the observation that information per-

taining the issues on the left of Figure 1 arrive at a much faster rate than the information concerning

the issues on the right of the figure. The idea is then to use this property of the problem to identify

an integrated way of approaching these issues that is self-consistent, and transparent. The primary

objective is not to produce more accurate numbers by incorporating intraday variations, but rather

define the representational structures that will allow qualitative and quantitative reasoning about the

effects of different time scales to take place. At the same time this structure can be used to perform

a dimensionality reduction on the original model and therefore allowing useful computation to be

performed.

Figure 2(a) shows a pictorial representation of this research programme. Given two models M1 and

M2 that have different dynamics, the aim is to construct bridges along the spatial, temporal and state

dynamics of the two models. Figure 2(b) shows a different approach to integrated modeling where

each model has its own dynamics and the two models interact when needed, e.g. when the output of

one model is used as the input to the other. The proposed approach uses the conceptual framework of

figure 2(b). Both approaches offer viable solutions to integrated modeling. In order to be specific we

only discuss two case studies. These two problems provide a partial implementation of the programme

outlined above. The first case study is concerned with the coupling of a capacity expansion model with

a unit commitment model. The question addressed here is how to incorporate hourly dynamics in the

capacity expansion model. The latter model is mainly concerned with investment decisions that are

6

Time

State

Space

M1 M2

(a) Integrated multiscale model-ing

Time

State

Space

M1

Time

State

Space

M2

(b) Integrated modelling

Figure 2: Two approaches on integrated modelling.

taken only every few years and their effects are long lived. The hourly dynamics describe electricity

load demand and the intermittency of wind generation. This case study is thus concerned with the

large scale integration of renewables in an existing portfolio of power generation. The objective is to

show how to link the two models along their state dimension.

The second case study addresses some of the challenges when considering the large scale integration

of distributed generation or microgrids. This problem is addressed from the point of view of a system

operator that may be concerned with assessing the impact of various technologies or deciding the

optimal size of a microgrid. This problem has a spatial aspect that is not present in the first case

study and so we use it as a template to illustrate how one could approach problems with a strong

spatial component using a multiscale framework.

3.1 Literature Review

3.1.1 Literature related to Case Study I

Both deterministic and stochastic models exist that address different aspects of the capacity expansion

problem in the electricity industry. Capacity expansion problems have generated a large amount of

literature. This is mainly because expansion problems are applicable to many areas but also because

they are a good testbed for new modeling ideas. We will only discuss models that address the effect

of the different time scales. Even though some of the work discussed below does not address capacity

expansion directly, we believe that the most relevant papers to this work are the ones that address

multiple scales since their ideas could be used in the capacity planning problem.

In [Sen et al., 2006] the problem of incorporating different scales for addressing risk management

7

problems such as buying and selling forwards for fuel are addressed in conjunction with the intraday

unit commitment problem. Their model does not address intraday effects from intermittent sources.

They formulate the problem as a multistage stochastic programming problem. The resulting large scale

mixed integer linear programming problem is solved using a nested column decomposition algorithm.

In [Epe et al., 2009] the authors do address the problem of dealing with the intermittency of wind.

Again a multistage stochastic programming approach is taken and the resulting large scale optimization

problem is solved using a recombining tree methodology. In [Pritchard et al., 2005] the operation of a

hydro-electric reservoir is addressed. Their problem also has multiple scales since the supply of power

occurs intraday, in hourly intervals, but the management of the reservoir occurs over monthly scales.

The problem is formulated as a dynamic programming problem. By approximating the decision to

have some desirable properties the different scales can be decomposed. In [Powell et al., 2009] the

problem of multiple scales as well as the intermittency of wind and solar are handled using stochastic

dynamic programming. In order to solve the intractable dynamic programming problem approximate

dynamic programming is used. We refer to the review article in [Wallace and Fleten, 2003] and the

recent book [Weber, 2005] for a more complete overview of both stochastic programming approaches

as well as approaches based on dynamic programming. What all these papers have in common is that

they address the existence of multiple scales using some sort of algorithmic framework. They either

use a decomposition algorithm, approximate dynamic programming, or find some other way to relax

the non-anticipativity constraints in order to make the problem tractable.

To make operational the uncertainty structures present in this class of problems we make use of

the tools from singular perturbation theory for MDPs in continuous time. In some respects some

models already take advantage of this structure. For example the widely used MARKAL model

[Seebregts et al., 2001] uses the concept of a “typical” load to overcome the onerous requirement to

optimize over all possible loads. This type of heuristic is useful, but as we are trying to develop models

to understand the behavior of systems we do not observe it is important to understand why heuristics

work, when they fail, and what can be done instead. For example it is not clear how to extend these

heuristics to handle wind intermittency, or demand elasticity (a major objective of demand response

programs).

The main results for multiscale MDPs are summarized in two excellent books [Sethi and Zhang, 1994,

8

Yin and Zhang, 1998]. Therefore we only comment on the relations between this work and the lit-

erature. In [Jiang and Sethi, 1991] a similar model to ours is proposed. The model is motivated

by a manufacturing system with machines that have failure rates that occur on fast time scales.

However, their model does not address the issue of capacity expansion and demand is also determin-

istic. In [Zhang et al., 1997] another model similar to ours is proposed and is again studied in the

context of manufacturing systems. Their model also has fixed capacity and deterministic demand.

Even though the problem of capacity expansion with multiple scales has been studied, see Ch 10

in [Jiang and Sethi, 1991], their model does not address capacity expansion for Markov Chains with

weak interactions and multiple scales. Moreover all the papers described above are concerned mainly

with manufacturing systems, and in order to carry the results over to the power system sector we

had to devise a new framework. This is because in order to capture some of the basic features of the

energy planning models we needed a model that can handle a Markov chain with weak interactions

both on capacity (to model wind intermittency for example) as well as on the demand dynamics (to

model the demand load dynamics). We also need to handle uncertainties that are not changing fast,

such as the long term demand growth for electricity. In Section 4 we only provide an outline of these

extensions. The details can be found in [Parpas and Webster, 2010].

3.1.2 Literature related to Case Study II

The literature on network planning for power systems is vast. We refer the interested reader to

[Wallace and Fleten, 2003] and [Georgiadis et al., 2008] for a review. However there are currently

no studies that attempt to address the problem of optimal planning in the context of large scale

penetration of distributed generation technologies. There are a few studies that discuss the prob-

lem of a single network [Hawkes and Leach, 2009, Abu-Sharkh et al., 2006]. Some heuristic methods

for this problem are discussed in [Mendez et al., 2006]. In our analysis we rely on the results from

[Chow and Kokotovic, 1985]. The model discussed in [Chow and Kokotovic, 1985] pertains issues at a

smaller scale than the ones we propose to study. Also the work in [Chow and Kokotovic, 1985] has no

controls and is concerned with the deterministic case. We are interested in stochastic optimal control

problems.

9

4 Case Study I: State aggregation for capacity expansion

Before we delve into the details of the model, we first motivate our construction by looking at some of

the characteristics of some real electricity demand data. Hourly electricity load data time series have

a strong periodic component. By far the most widely used method to analyze such data sets is a two

step process. The first step involves identifying the periodic component in order to be able to remove

the trends that are related to time of day, weekend, and any other purely temporal elements. The

second step is to fit some stochastic process to the remaining component (we call this the stochastic

component). Hourly load data for 2007 were used from the PJM midatlantic region (PJM.E). This

data is freely downloadable from the PJM website.

Electricity hourly load demand is modeled as:

H(t) = D(t) + η0(t), t ∈ [0, T ]. (4.1)

The deterministic component is given by,

D(t) = a + bt +

N∑

j=0

cj cos(2πφjt + lj) t ∈ [0, T ]. (4.2)

Thus the deterministic component accounts for a linear trend in demand. The seasonal fluctuations

are super imposed around this linear trend. In our numerical experiments we used N = 5. The

deterministic harmonics of (4.2) are obtained from the peaks of the Fourier transform of the hourly

load data. In a more realistic model these parameters would be uncertain. The term η0 represents

the stochastic component. In order to identify a suitable stochastic process for η0 we fit a maximum

likelihood Markov Chain to the residual component. The residual component is given by,

r(t) = A(t) − D(t),

where A(t) represents the actual data (8760 points). The resulting transition matrix of the Markov

Chain is shown in Figure 3. The states in the Markov Chain represent the amount by which the

hourly load is above or below the deterministic periodic component (for ease of exposition the axis

shows the state number rather than the actual state). What is immediately obvious from this figure

10

is that apart from a few cases the Markov Chain will tend to stay around the same state. Of course

the variations around the state, as well as far away from the diagonal are important since they will

influence the amount of reserves. Figure 3(b) shows the result of the same process but with using

a Markov Chain with only ten states. Using typical states, optimizing over these and summing the

results, essentially assumes a transition matrix as in Figure 3(c) (an identity matrix). However using

typical states will always underestimate the costs of running such a system (this follows from Jensen’s

inequality and will be true for convex models, such as the one studied in this paper). This means that

a certain amount of guesswork will always be required in order to find the level of reserves, or any

other quantity of interest. Therefore, it seems that assuming the problem to be reducible to a few

typical states oversimplifies the problem. In view of the drive towards more efficient management of

increasingly scarce resources it is important to attempt to find better ways of incorporating hourly load

demand that leads to more realistic yet tractable models. Introducing uncertainty into demand load

management will give us a better understanding both of the true costs of the system as well as help

identify strategies to manage the dynamic equilibrium between supply and demand. At the same time

it is obvious from the figure that the data does not completely lack structure. To see the difference, a

process with no structure is shown in Figure 3(d). Therefore assuming a completely general structure

is also not appropriate. Methods that do not take advantage of this highly specialized feature of the

problem are essentially implicitly assuming a structure such as the one shown in Figure 3(d).

We followed a similar procedure for the wind data. We obtained wind data from the NCDC

[NCDC, 2009] website. The wind data contains the wind speeds at an hourly time interval over the

years 2007–2008, in the Buzzard’s bay area in south Massachusetts. The resulting Markov Chain for

wind uncertainty is shown in Figure 4.

The information structure proposed in this paper is shown in Figure 1. We start from the

bottom layer that has all the detailed intraday variations. This information is combined into ag-

gregate states. Finally some other probabilities are devised that control the transitions between

aggregate states. In order to make the structure in Figure 1 operational, we deploy the tools of

singular perturbation theory. Below we describe the main components of this framework. More infor-

mation can be found in [Yin and Zhang, 1998, Zhang et al., 1997, Phillips and Kokotovic, 1981] and

[Parpas and Webster, 2010].

11

(a) A Markov chain with 200 states (b) A Markov chain with 10 states

(c) (d) A structure free transition matrix

Figure 3: Markov transition matrices. (a)-(b) are obtained from a maximum likelihood fit to hourlytime series. (c) is the structure implicitly assumed when considering typical states. (d) was obtainedby a maximum likelihood fit to a white noise process.

12

(a) 20 states (b) 50 states

Figure 4: Markov transition matrices for wind uncertainty. (a)-(b) are obtained from a maximumlikelihood fit to hourly time series

Motivated by the data above we will assume a specific structure for the uncertainties representing

load demand and wind availability. To this end, the transition function is assumed to satisfy:

dP

dτ= p(τ)[Q + ǫW ],

p(0) = I.

(4.3)

where Q = diag(Q1, . . . , Qm). For the purposes of this section, we assume that the m aggregate states

have been identified by some statistical procedure. The ǫ parameter is used to make explicit the

assumption that the transitions inside one of the aggregate states are much more common than the

transitions between aggregate states. Since ǫ is assumed to be small, if we run the system for a small

amount of time then the W matrix will have a small role to play. However, if we run the system for

long periods then the transitions between aggregate states become significant. In order to capture the

latter effects, even as ǫ approaches 0, we stretch the time dimension t = ǫτ . This stretching of time

is what sets singular perturbations apart from regular perturbations. With this change of time (4.3)

13

becomes:

dP

dt= p(t)[

1

ǫQ + W ],

p(0) = I.

We note that we have not made any approximations so far. All we have assumed is that we can

identify the different aggregate states, and that we can specify an ǫ > 0 as a small parameter. This

parameter is used as a measure of the separation of scales. We are now in a position to fully specify

our model.

4.1 The model

We assume that the decision maker can use any of the N generating plants to meet demand

dx(t)

dt=

N∑

i

ui(t) − z(t) + η0(t, ǫ) x(0) = x. (4.4)

Where ui represents the output from the ith plant. x(t) represents the amount of energy not served.

We assume that some suitable penalty is imposed for meeting demand i.e. the cost of each unit of x

is high compared to the cost of production. z(t) represents the demand, and it is assumed to have the

following dynamics,

dz(t)

dt= f(z(t), ξ0(t)) z(0) = z. (4.5)

In (4.4) η0 is the fast scale Markov process that represents the intraday variations in hourly loads.

In (4.5) ξ0 is also assumed to be a finite state Markov process, but this process changes at a slower

rate. The latter process is used to represent uncertainties such as the long term trend demand for

electricity. Information about long term demand arrives at longer time intervals. Therefore it is not

appropriate to model this type of uncertainty as a fast scale process. We use the notation η and ξ, for

the fast, and slow scale dynamics throughout the paper.

We also assume that the decision maker can invest in each of the generating facilities with the

aim of increasing capacity. Some of the plants are allowed to have zero capacity to start with. The

14

cumulative investment, in the ith plant, is given by:

dyi(t)

dt= πi(t) yi(0) = yi, i = 1, . . . , N.

By imposing appropriate upper bounds on the rate of investment, πi, we can ensure that new capacity

cannot appear overnight. We assume that once total investment in plant i reaches level Ki then new

capacity becomes available. Each plant can be expanded once, and the level of investment required is

assumed to be deterministic. Both of these assumptions can be relaxed, but with significant increase

in the complexity of the analysis.

The total available capacity in the ith plant is assumed to be a Markov process, and it is given by:

ηi(ǫ, t) =

ηi1(ǫ, t) t ≤ τi

ηi2(ǫ, t − τi) t > τi

Where τi represents the stopping time:

τi = inf{t | yi(t) = Ki} ∧ T.

The states of ηi1(ǫ, t) and ηi

2(ǫ, t) are given by the set Si1 and Si

2 respectively. Where

Si1 = {si

1,1, si1,2, . . . , s

i1,mi

},

similar notation is used for Si2. We also assume that there exists a one-to-one mapping between states

in Si1 and Si

2, this will be denoted by κ : Si1 → Si

2. We use the notation Si to denote the states of

available capacity in the ith plant.

The Markov process η(t, ǫ) = [η0(t, ǫ), . . . , ηN (t, ǫ)] is assumed to have a generator given by:

dpη(t)

dt= pη(t)[W +

1

ǫQ] (4.6)

Where Q = diag(Q1, . . . , Ql). Furthermore both W and {Qi}li=1 are assumed to be irreducible. The

generator for the ξ variables can be much more general since we will not perform any approximation

15

on this process. The dynamics of ξ are given by,

dpξ(t)

dt= pξ(t)R.

A policy (u(t), π(t)) will be called admissible if it satisfies the dynamics specified above (including

the stochastic bounds on available capacity). Moreover an admissible policy needs to be adapted to

the filtration Ft. The set of all Ft-adapted processes is denoted by Aǫt. For notational simplification

we use w(t) = [x(t), y(t), z(t)] wherever possible.

The objective function and the complete problem are given below.

Jǫs(w, η, ξ;u, π) = E

{N∑

i=1

∫ T

s

e−ρ(t−s)Gi(xs, πi(s), ui(s), ξ(s))ds + e−ρ(T−s)φ(XT )

}

vs(w, η, ξ) = min Jǫs(w, η, ξ;u, π)

dx(t)

dt=

N∑

i

ui(t) − z(t) + η0(t, ǫ) x(s) = x

dz(t)

dt= f(z(t), ξ0(t)) z(s) = z

dy(t)

dt= π(t), y(s) = y (Hǫ)

yi(t) ≤ Ki

0 ≤ ui(t) ≤ ηi(t, ǫ)

(u(t), π(t)) ∈ Aǫt

Where Aǫt denotes the set of admissible controls. Note that the objective is allowed to depend on the

slow Markov process ξ. This may be used to represent uncertainties in operating costs (e.g. fuel costs,

emission costs etc.).

16

4.2 Asymptotic Analysis

As ǫ goes to zero, it can be shown [Parpas and Webster, 2010] that the value function of (Hǫ) is the

solution of the following aggregate problem.

vs(w, k, ξ) = min Js(w, k, ξ;u, π)

dx(t)

dt=

j∈Mη(t)

λη(t)j

( N∑

i

uη(t),ji (s) − z(t) + η

η(t),j0

)x(s) = x

dz(t)

dt= f(z(t), ξ0(t)) z(s) = z

dy(t)

dt= π(t), y(s) = y (H)

yi(t) ≤ Ki

uη(t),ji (t) ∈ Ci(η(t), j)

η(s) = k.

Note that the value function is not a function of η but rather of the current aggregate state the process

is in. Ck(i, j) is the feasible set associated with the ith state inside the jth aggregate set of states for

technology k. In the case with no startup costs then this will simply be given by:

0 ≤ uη(t),ji (t) ≤ η

η(t),ji .

The expectation is taken with respect to the aggregated Markov Chain. That is the Markov Chain

that specifies the probability of moving between clusters of states. The complete derivation of the

approximate problem can be found in [Parpas and Webster, 2010]. The important feature in (H)

is that we only consider aggregate sets of states. This has significant computational advantages.

Additional numerical experiments from [Parpas and Webster, 2010] show that the algorithm used to

solve (H) requires a lot less iterations to converge than when the same algorithm is used on the

original problem. This is because of the so called smoothing property for such problems. We refer the

interested reader to [Parpas and Webster, 2010] for the details.

17

5 Case Study II: Spatial aggregation for distributed generation plan-

ning

Traditionally the physical flow of electricity has been fixed and centrally managed. Large generating

plants (hydro, nuclear or fossil-fired) generate power, and through the high voltage transmission net-

work, it reaches the low voltage network and eventually gets distributed to consumers. The economics

of power generation made this mode of operation efficient. Deregulation, concerns over environmental

impacts of power generation and security issues are challenging this traditional structure of power

generation. Distributed generation is the term used to describe a set of technologies that are located

close to the load demand centers. Since these distributed resources are not centrally managed they

pose a fresh set of challenges for the modeling of power grids. The term microgrid is used to de-

scribe a set of decentralized technologies operating to provide heat and power to a small set of mostly

residential customers. The technologies used in microgrids include both renewables sources such as

solar, wind, but also fossil-fuel based generators. A distinctive feature of microgrids is the empha-

sis on polygeneration. A microgrid can operate in isolation to the large national grid. It may also

operate in tandem with the main grid. The tandem operation will be needed if the microgrid can

not generate enough local power to satisfy local demand, or when too much power is generated and

then it can be sold back to the main grid. More information on the engineering aspects of distributed

generation can be found in [Masters, 2004]. Most studies have looked at the impact of a single mi-

crogrid (e.g. [Hawkes and Leach, 2009, Abu-Sharkh et al., 2006]). With some notable exceptions (see

[Mendez et al., 2006]) the large scale integration of microgrids has not been sufficiently analyzed. The

aim of this section is to show how a large number of microgrids can be analyzed together using similar

techniques as in Section 4.

5.1 The model

The first step in extending the techniques of section 4 is to endow the Markov process, representing the

load demand or any other uncertain quantity such as the availability of wind, with a spatial structure.

This class of stochastic processes are called random fields, see e.g. [Whittle, 1986]. Random fields

have been studied extensively. The optimal control of Markov random fields has received very little

attention compared with the standard Markov decision processes. Most of the work on the control of

18

spatial Markov processes is reviewed in [Chornei et al., 2006].

Let Γ = (V,Be, Bp) denote an undirected graph, where V denotes the set of vertices, and Bp

the set of physical edges, and Be the electrical edges. A physical edge exists between each node if

there is conditional dependency (in the sense defined below) between two nodes. An electrical edge

between two nodes means that a node is electrically connected with another node. Obviously this

is an idealization of how a large electrical network might be connected, however when attempting to

incorporate large amounts of intermittent wind generation it becomes important to separate the two

concepts. The graph is illustrated in Figure 5. The electrical graph should be easy to identify given

the actual configuration of the network. The graph related to the physical aspects of the network may

be different, even though one would expect that nodes that are close to each other should belong to the

same neighborhood. Each node on the graph will represent a single source of distributed generation.

Depending on the application this may be down to a single microgrid or a cluster of microgrids.

By k ↔ j we denote that vertex j is connected to vertex k. For each k ∈ Bp we define the set

Np(k) = {j ∈ Bp | k ↔ j} as the set of all neighboring vertices to vertex k. We use Np(k) = Np(k)∪k.

We use similar notation for Be. We further assume that the graph can be split into l different areas.

We use Mk to denote the set of nodes in area k.

At each node on the graph we use cIi (cE

i ) to denote the number of internal (external) connections

at node i. We define,

cI = mini{cI

i } cE = maxi

{cEi } ǫ1 =

cE

cI.

Therefore as ǫ1 gets smaller, one would expect to be able to analyze each area of the graph separately

without much loss of information. Next we define sparsity parameters for each area. We use sIk (sE

k )

to denote the number of internal (external) connections in area k. We use m = mink{|Mk|}. As in

the case of the node sparsity parameter we also define,

sI = mink

{sIk} sE = max

k{sE

k } ǫ2 =sE

mcI.

Intuitively, one would expect to be able to analyze each area separately when ǫ1 and ǫ2 are small

enough. Similar techniques as in the previous section could be used. In what follows we formulate a

simple network in order to make this operational.

19

The stochastic process ηi(t, ǫ) will be used to represent the random variables at vertex i. We will

assume the following,

P [ηi(τ, ǫ) = η | Ft] = P [ηi(τ, ǫ) = η | ηj(t, ǫ), j ∈ Np(i)], τ > t, (5.1)

where Ft =⊗

k∈V Fk. It is clear that the preceding equation is a generalization of the familiar

Markov condition to the case when the problem has spatial structure. See [Whittle, 1986] for more

details regrading the properties of such processes. It is natural to expect this type of processes to be

useful for the modeling of wind availability. Demand uncertainty can be assumed to be conditionally

independent from its neighbors (after the temporal component has been eliminated by the methods

described earlier). Consider the following equations describing the state of vertex i in area k

(a) Physical Graph (b) Electrical Graph

Figure 5: The two types of graphs representing the spatial structure of the model.

dxki

dt= rǫ

k,i(t) +∑

j∈Ne(i)

Lij , (5.2)

where,

rǫk,i(t) =

ni∑

j=1

ujk,i(t) − η

i,k0 (t, ǫ),

represents the power produced locally by the ith microgrid. As before u represent the power output,

and each node is assumed to have an ni number of generators. Lij represents the net injection of power

between nodes i and j. In general Lij would be a control. However in order to make the ensuing

20

analysis tractable we will assume that,

Lij = Lij(xi − xj).

Note that this function must satisfy an energy balance condition,

i

j∈Ne(i)

Lij = 0.

To help make the analysis concrete we assume the following,

Lij =

λij(xj − xi) if i ↔ j

0 otherwise

The general case of nonlinear functions can be handled in a similar way as in [Chow, 1982]. Under

these assumptions the system in (5.2) can now be written as,

dxki

dt= rǫ

k,i(t) +∑

j∈Mk

j 6=i

λij(xkj − xk

i ) +

l∑

m6=k

j∈Mm

λij(xkj − xk

i ). (5.3)

Motivated by the decomposition above, we define LI ∈ Rn×n to be a matrix containing the internal

connections in an area. This implies that LI must be block diagonal. The diagonal entries in this

matrix are chosen to satisfy,

LIii = −

j∈Mk

λij ∀i ∈ Mk ∀k.

The off-diagonal entries of LI are chosen to equal λij provided that i and j belong to the same group.

Similarly we define LE ∈ Rn×n to be a matrix containing external connections. Thus LE satisfies the

following two conditions,

ΛEij =

λij if i 6= j and i ∈ Mk, j ∈ Mm m 6= k

−∑

m6=k

∑j∈Mm

λij if i = j, i ∈ Mk

0 otherwise.

21

Note that by definition L = LI + LE . With the definitions given above we can decompose the system

into its external and internal connections.

dx

dt= rǫ(t) + (LI + LE)x(t), . (5.4)

Define the following aggregate variable,

yk =

|Mk|∑

i=1

xki

|Mk|=

1

|Mk|1T|Mk|

xk,

where 1n denotes a vector of ones of dimension n. We can vectorize the expression above as follows,

y = Cx.

Where,

C , M−1UT ∈ Rl×n,

U , diag(1|M1|, . . . ,1|Ml|) ∈ Rn×n,

M , diag(|M1|, . . . , |Ml|) ∈ Rl×l.

Next we define the first node in each neighborhood as a reference node,

zki−1 = xk

i − xki−1 i = 2, . . . ,Mk.

We will use z = Gx, to denote the equation above in vector form. The matrix G is defined in the

obvious way. It is shown in the Appendix that after some change of variables the following scaled

system can be derived,

dy

dτ= G1r

ǫ(t) + A1y + A2z

ǫ2dz

dτ= G1r

ǫ(t) + ǫ1B1y + B2z.

It can be seen from the system above that ǫ2 is a singular perturbation parameter, while the parameter

ǫ1 serves the role of a regular perturbation parameter. The significance of the system in 5.1 is that it

22

represents a controlled approximation of the aggregated dynamics of the original system. As there are

three small parameters for this problem1, the asymptotic analysis as these parameters tend to zero is

much more complicated. Despite the increased complexity there are several ways one could proceed.

One could solve the optimal control problem by letting ǫ approach to zero. At this limit the dynamics

representing the uncertainties in net power flow in each area would be obtained. The solution will

contain two sources of error. The first is due to taking the limit ǫ → 0. The second source of error is

due to using the aggregated dynamics instead of the correct dynamics. Error estimates for the latter

source of error could then be derived using standard techniques (see e.g. [Chow and Kokotovic, 1985]).

Due to space limitations we omit the detailed implementation of this approach. The details will appear

elsewhere.

6 Conclusions

Optimization models have been used in the electric power industry for a long time. Environmental

considerations and the development of new distributed generation technologies make it necessary to

consider the properties of power networks at multiple scales and from the point of view of multiple

agents. This paper proposed a tentative framework to address some of these challenges.

Appendix

In the Appendix we show how the perturbed dynamics in (5.1) are derived. These calculation were

done in [Chow and Kokotovic, 1985]. One difference in the computations below is the existence of

control variable representing the net injection of power in a node. The other difference is the existence

of uncertainty. Using the same notation as in Section 5, we obtained the following two expressions for

the nodes on the graph,

y = Cx z = Gx,

It follows from [Aoki, 1968] that we can derive the following expression for x,

x = Uy + Wz,

1Remember that ǫ is the singular perturbation parameter for the uncertainties around net power injection.

23

where W = GT (GGT )−1. It follows that,

dy

dt= C

dx

dt= C(rǫ(t) + (LI + LE)x(t))

= C(rǫ(t) + (LI + LE)(Uy(t) + Wz(t)))

= Crǫ(t) + CLEUy(t) + CLEWz(t).

(6.1)

Similarly, for z we obtain the following,

dz

dt= G

dx

dt= G(rǫ(t) + (LI + LE)x(t))

= Grǫ(t) + GLEUy(t) + (GLEW + GLIW )z(t).

(6.2)

The significance of the reformulation of the system above is that the matrix with the internal links

LI appears only in the last expression in the system above. Next we observe that if ǫ1 and ǫ2 are

sufficiently small2 then one would expect the norm of (GLEW + GLIW ) to be much larger than the

other matrixes in (6.1) and (6.2). This motivates the following change of time t1 = ǫ1t. We also rescale

the matrices to that everything is of O(1). We use the following definitions,

A1 ,CLEU

cIǫ2A2 ,

CLEW

cIǫ2

B1 ,GLEU

cIǫ1B2 ,

G(LI + LE)W

cI.

G1 ,G

cI

With the rescaled dynamics introduced above we obtain the following dynamics,

dx

dt1= ǫ2G1r

ǫ(τ) + ǫ2A1y + ǫ2A2z

dz

dt1= G1r

ǫ(τ) + ǫ1B1y + B2z.

The final system is obtained by rescaling the system above in its slow time scale,

τ = ǫ2t1

2Remember that ǫ1 is the perturbation parameter representing the ratio of external to internal connection in aneighborhood. The other parameter ǫ2 represents the permutation parameter that represents the sparsity of the graphin relation to the number of nodes and number of internal connections.

24

dx

dτ= G1r

ǫ(t) + A1y + A2z

ǫ2dz

dτ= G1r

ǫ(t) + ǫ1B1y + B2z.

References

[Abu-Sharkh et al., 2006] Abu-Sharkh, S., Arnold, R., Kohler, J., Li, R., Markvart, T., Ross, J.,

Steemers, K., Wilson, P., and Yao, R. (2006). Can microgrids make a major contribution to UK

energy supply? Renewable and Sustainable Energy Reviews, 10(2):78–127.

[Aoki, 1968] Aoki, M. (1968). Control of large-scale dynamic systems by aggregation. IEEE Trans-

actions on Automatic Control, 13(3):246–253.

[Bloom, 2009] Bloom, J. (2009). Optimization Applications in the Energy and Power Industries.

http://www-01.ibm.com/software/websphere/industries/energy/.

[Bohringer and Rutherford, 2008] Bohringer, C. and Rutherford, T. (2008). Combining bottom-up

and top-down. Energy Economics, 30(2):574–596.

[Cash and Moser, 2000] Cash, D. and Moser, S. (2000). Linking global and local scales: designing

dynamic assessment and management processes. Global Environmental Change, 10(2):109–120.

[Chornei et al., 2006] Chornei, R., Daduna, H., and Knopov, P. (2006). Control of spatially structured

random processes and random fields with applications. Springer Verlag.

[Chow, 1982] Chow, J. (1982). Time-scale modeling of dynamic networks with applications to power

systems. Springer-Verlag.

[Chow and Kokotovic, 1985] Chow, J. and Kokotovic, P. (1985). Time scale modeling of sparse dy-

namic networks. IEEE Transactions on Automatic Control, 30(8):714–722.

[Claudio et al., 2008] Claudio, C., Gomez-Exposito, A., Conejo, A., Canizares, C., and Antonio, G.

(2008). Electric Energy Systems: Analysis and Operation. CRC.

[Epe et al., 2009] Epe, A., Kuchler, C., Romisch, W., Vigerske, S., Wagner, H., Weber, C., and Woll,

O. (2009). Optimization of Dispersed Energy Supply-Stochastic Programming with Recombining

Scenario Trees. Optimization in the Energy Industry, page 347.

25

[Georgiadis et al., 2008] Georgiadis, M., Pistikopoulos, E., and Kikkinides, E. (2008). Energy Systems

Engineering: Volume 5: Energy Systems Engineering. Wiley-VCH.

[Hawkes and Leach, 2009] Hawkes, A. and Leach, M. (2009). Modelling high level system design and

unit commitment for a microgrid. Applied Energy, 86(7-8):1253–1265.

[Jiang and Sethi, 1991] Jiang, J. and Sethi, S. (1991). A state aggregation approach to manufacturing

systems having machine states with weak and strong interactions. Operations Research, 39(6):970–

978.

[Masters, 2004] Masters, G. (2004). Renewable and efficient electric power systems. Wiley, Inter-

science.

[Mendez et al., 2006] Mendez, V., Rivier, J., Fuente, J., Gomez, T., Arceluz, J., Marin, J., and

Madurga, A. (2006). Impact of distributed generation on distribution investment deferral. In-

ternational Journal of Electrical Power and Energy Systems, 28(4):244–252.

[NCDC, 2009] NCDC (2009). National Climatic Data Center,

http://www.ncdc.noaa.gov/oa/ncdc.html.

[Parpas and Webster, 2010] Parpas, P. and Webster, M. (2010). A singular perturbation approach to

multiscale power system capacity expansion. Submitted.

[Parsons, 2008] Parsons, J. (2008). Do Trading and Power Operations Mix? The Case of Constellation

Energy Group 2008. Massachusetts Institute of Technology, Working Papers, Center for Energy and

Environmental Policy Research.

[Phillips and Kokotovic, 1981] Phillips, R. and Kokotovic, P. (1981). A singular perturbation ap-

proach to modeling and control of Markov chains. IEEE Transactions on Automatic Control,

26(5):1087–1094.

[Powell et al., 2009] Powell, W., George, A., Lamont, A., and Stewart, J. (2009). SMART: A Stochas-

tic Multiscale Model for the Analysis of Energy Resources, Technology and Policy. Preprint.

[Pritchard et al., 2005] Pritchard, G., Philpott, A. B., and Neame, P. J. (2005). Hydroelectric reservoir

optimization in a pool market. Math. Program., 103(3, Ser. A):445–461.

26

[Seebregts et al., 2001] Seebregts, A., Goldstein, G., and Smekens, K. (2001). Energy/environmental

modeling with the MARKAL family of models. In Proc. Int. Conf. on Operations Research (OR

2001).

[Sen et al., 2006] Sen, S., Yu, L., and Genc, T. (2006). A stochastic programming approach to power

portfolio optimization. Oper. Res., 54(1):55–72.

[Sethi and Zhang, 1994] Sethi, S. P. and Zhang, Q. (1994). Hierarchical decision making in stochastic

manufacturing systems. Systems & Control: Foundations & Applications. Birkhauser Boston Inc.,

Boston, MA.

[Wallace and Fleten, 2003] Wallace, S. and Fleten, S. (2003). Stochastic programming in energy. In

A., R. and A., S., editors, Handbooks in OR & MS, chapter 10. Elsevier Science.

[Weber, 2005] Weber, C. (2005). Uncertainty in the electric power industry: methods and models for

decision support. Springer Verlag.

[Whittle, 1986] Whittle, P. (1986). Systems in stochastic equilibrium. John Wiley & Sons, Inc. New

York, NY, USA.

[Wing, 2006] Wing, I. (2006). The synthesis of bottom-up and top-down approaches to climate policy

modeling: Electric power technologies and the cost of limiting US CO2 emissions. Energy Policy,

34(18):3847–3869.

[Yin and Zhang, 1998] Yin, G. G. and Zhang, Q. (1998). Continuous-time Markov chains and ap-

plications, volume 37 of Applications of Mathematics (New York). Springer-Verlag, New York. A

singular perturbation approach.

[Zhang et al., 1997] Zhang, Q., Yin, G., and Boukas, E. (1997). Controlled Markov chains with weak

and strong interactions: asymptotic optimality and applications to manufacturing. Journal of

Optimization Theory and Applications, 94(1):169–194.

27