Sie sind auf Seite 1von 17

PROGRESS IN PHOTOVOLTAICS: RESEARCH AND APPLICATIONS Prog. Photovolt: Res. Appl.

(2011) Published online in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/pip.1204

PAPER PRESENTED AT 26TH EU PVSEC, HAMBURG, GERMANY 2011

Is the distribution grid ready to accept large-scale photovoltaic deployment? State of the art, progress, and future prospects
Martin Braun1,2*, Thomas Stetz2, Roland Brndlinger3, Christoph Mayr3, Kazuhiko Ogimoto4, Hiroyuki Hatta5, Hiromu Kobayashi5, Ben Kroposki6, Barry Mather6, Michael Coddington6, Kevin Lynn7, Giorgio Graditi8, Achim Woyte9 and Iain MacGill10
1 2

IEH, University of Stuttgart, Pfaffenwaldring 47, 70569 Stuttgart, Germany Fraunhofer Institute for Wind Energy and Energy System Technology (IWES), Koenigstor 59, 34119 Kassel, Germany 3 Austrian Institute of Technology (AIT), Vienna, Austria 4 Institute of Industrial Science (IIS), University of Tokyo, Tokyo, Japan 5 Central Research Institute of Electric Power Industry (CRIEPI), Tokyo, Japan 6 National Renewable Energy Laboratory (NREL), Golden, CO, USA 7 US Department of Energy, Washington, DC, USA 8 ENEAItalian National Agency for New Technologies, Energy and Sustainable Economic Development, Naples, Italy 9 3E sa, Brussels, Belgium 10 Centre for Energy and Environmental Markets (CEEM), University of NSW, Sydney, Australia

ABSTRACT
The installed capacity of photovoltaic (PV) systems has recently increased at a much faster rate than the development of grid codes to effectively and efciently manage high penetrations of PV within the distribution system. In a number of countries, PV penetrations in some regions are now raising growing concerns regarding integration. Management strategies vary considerably by countrysome still have an approach that PV systems should behave as passive as possible, whereas others demand an active participation in grid control. This variety of grid codes also causes challenges in learning from best practice. This paper provides a review of current grid codes in some countries with high PV penetrations. In addition, the paper presents a number of country-specic case studies on different approaches for improved integration of PV systems in the distribution grid. In particular, we consider integration approaches using active and reactive power control that can reduce or defer expensive grid reinforcement while supporting higher PV penetrations. Copyright 2011 John Wiley & Sons, Ltd.
KEYWORDS photovoltaic; grid integration; distribution grid; grid codes; case studies; ancillary services *Correspondence M. Braun, Fraunhofer Institute for Wind Energy and Energy System Technology (IWES), Koenigstor 59, 34119 Kassel, Germany. E-mail: martin.braun@iwes.fraunhofer.de Received 15 May 2011; Revised 10 July 2011; Accepted 26 August 2011

1. INTRODUCTION
In 2005, the globally installed photovoltaic (PV) capacity was about 5.4 GW [1]. Driven by signicant policy incentives in some countries and cost reductions in PV systems, the annual global PV installation rate has since rocketed, leading to a worldwide installed PV capacity approaching 40 GW in 2010 [1]. In this period of extraordinary industry growth, some national markets have emerged as global leaders for the installation of new PV systems. Some expert assessments estimate that this industry growth may continue to grow over the coming years as shown in Table I.
Copyright 2011 John Wiley & Sons, Ltd.

This paper is a joint work of Germany, Austria, Japan, USA, Italy, Belgium, and Australia that represent a total market share in 2010 of more than 70%. In contrast to conventional power plants and wind energy converters, PV systems are typically connected to low-voltage (LV) and medium-voltage (MV) systems. In part this reects the inherent scalability of PV technologies (grid-connected system sizes can vary from hundreds of watts to hundreds of megawatts) and its ready integration into the built environment. It also reects the nature of policy support in some countries, which emphasizes distributed applications. In Germany, for example, up to 80% of the

Distribution grid and large-scale PV deployment

M. Braun et al.

Table I. Installed photovoltaic capacity in 2010 and the European Photovoltaic Industry Associations (EPIA) market outlook for 2015. 2010 Installed Market capacity (GW) share (%) 17.2 3.5 3.8 1.0 0.8 0.1 29.3 0.5 3.6 2.5 39.5 43.5 8.9 9.6 2.5 2.0 0.3 74.2 1.3 9.1 6.3 100 2015 Installed Market capacity (GW) share (%) 32.2 13.0 6.2 4.1 2.3 0.5 69.1 2.3 11.2 22 131 24.6 9.9 4.7 3.1 1.8 0.4 52.7 1.8 8.5 16.8 100

Country Germany Italy Spain France Belgium Austria EU Australia Japan USA Worldwide

The data for 2015 is based on EPIAs moderate scenario [1].

installed PV capacity is connected to LV systems, and almost all PV systems are connected to the distribution system [2]. This coupling characteristic leads to new challenges for a secure and reliable distribution system operation especially if one is taking into account that PV systems are generally spatially concentrated in some areas rather than being uniformly distributed for reasons including regional policy support measures and demographics. Typical technical problems in distribution systems related to a high local PV penetration are local overvoltages and equipment overloading. In countries with high penetrations such as Germany, this is increasingly leading to expensive grid reinforcement measures. Because of their capability to control their active and reactive power, however, PV inverters are well suited to provide different kinds of ancillary services in order to avoid or at least reduce grid reinforcement measures and to keep the overall costs for PV grid integration as low as possible. In an International Energy Agency research collaboration, these capabilities are analyzed in detail. It is presented in Section 2. Section 3 provides a review of present and prospective grid codes in some of the countries currently leading PV deployment. A number of country-specic case studies on different approaches for improved integration of PV systems in the distribution grid are then presented in Section 4. In particular, integration approaches using active and reactive power control are discussed, which allow reducing grid reinforcement costs.

to address this evolving eld and to collect and disseminate international knowledge of PV systems at high penetration levels, is becoming crucial for the further large-scale deployment of PV. Within the framework of the International Energy Agency Photovoltaic Power System program, the new Task 14 High Penetration of PV in Electricity Grids has a focus on the role of PV in electricity grid congurations with a high penetration of renewable energy sources. Although no common denition of high-penetration PV scenarios yet exists, there is common understanding that high PV penetration exists, if local additional efforts are necessary to maintain a secure and reliable distribution system operation. Task 14 analyzes particular issues related to a high penetration of PV in electricity grids in order to show the full potential of grid-connected PV systems. The main goal of Task 14 is to facilitate the use of gridconnected PV as an important source in electric power systems. Therefore, technical barriers have to be reduced to achieve high penetration levels of distributed renewable systems in the electric power system. Task 14 will develop technologies and methods that enable a widespread deployment of distributed PV technologies in the electricity grids at lowest overall costs. The work program of Task 14 addresses mainly technical issues, including energy management, grid interaction, and penetration aspects related to local distribution grids and central PV generation scenarios, including integration in buildings and focus on multifunctional inverters as a smart interface between PV and the electricity system.

3. GRID CONNECTION REQUIREMENTS


This section describes existing, recently modied and discussed grid codes for the connection of PV to the distribution grid for chosen countries: Germany, Japan, USA, Italy, Belgium, and Australia. Because technical guidelines are often complex documentations, the focus of the following country-specic examples is set on steadystate requirements and frequency control using active and reactive power control. 3.1. German grid codes In Germany, the voltage level of the PV systems point of common coupling (PCC) denes the technical requirements concerning active and reactive power control. 3.1.1. The German medium-voltage grid code The German technical guideline for the connection to the MV network [3] is valid since 2009. Some regulations, for example, reactive power control, were postponed up to April 2011 due to delays of commercially available PV inverters with required control functionalities.
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

2. INTERNATIONAL ENERGY AGENCY PHOTOVOLTAIC POWER SYSTEM RESEARCH COLLABORATION


With further growth of distributed as well as centralized PV capacities, the need for international research collaborations,

M. Braun et al.

Distribution grid and large-scale PV deployment

3.1.1.1. Steady-state voltage rise criteria. The voltage rise due to distributed generators must not exceed +2% of the nominal system voltage (usually 20 kVLL or 10 kVLL).

3.1.1.2. Active power control. The PV system must be capable of reducing its active power output in cases of danger for a proper system operation. These cases include the following: potential congestions and equipment overloading, imminent danger of islanding operation, endangerment for static or dynamic system stability, over frequency, and maintenance work.

3.1.1.3. Reactive power control. In general, PV systems connected to the MV network have to be capable of providing a minimum power factor of 0.95 leading and lagging. It is up to the local distribution system operator (DSO) to dene the method how reactive power should be provided. Table II gives an overview about possible reactive power provision methods and their properties. 3.1.2. The German low-voltage grid code The technical guideline for the connection to the LV network [4] has been revised in August 2011 and is described in the following. In contrast, the former grid code VDE 0126 did not dene any active or reactive power control requirements, and the steady-state voltage rise due to distributed generators is limited to +2%. 3.1.2.1. Steady-state voltage rise criteria. The voltage rise due to distributed generators must not exceed +3% of the nominal system voltage (400VLL). 3.1.2.2. Active power control. Photovoltaic systems connected to the LV network must be capable of operating with a reduced active power output. The cases where a reduction of the active power output by the DSO is permitted are the same as for PV systems at the MV level. In contrast to the MV guideline, there is no demand for providing a remote control interface, except for generators with more than 100-kW connection capacity (}9 EEG). In terms of frequency control, the same requirements apply as for generators connected to the MV level. Please note that by April 2011, an interim solution for the so-called 50.2-Hz problem was released [5]. From April 2011 on and until the release of the guideline, all generators either have to be equipped with the 40% droop function or with randomly assigned threshold values between 50.3 Hz and 51.5 Hz. 3.1.2.3. Reactive power control. According to the technical guideline, all PV systems must have the capability to provide reactive power. The minimum power factor depends on the maximum inverter apparent power of the PV system. Table III gives an overview of the different requirements.

Furthermore, active power reduction is exceptionally permitted in cases where local grid reinforcement measures, according to }9 German Renewable Energy Sources Act (EEG), are not yet completed. All generators, connected to the MV system have to be capable of receiving remote set values for their active power output. The applied remote control technology is not yet standardized. If a reduction of the active power output is required, the PV system must be capable of reducing its active power output immediately, with single reduction steps not exceeding 10% of the PV connection capacity. Recommended set values for the operation of the PV system are 100%, 60%, 30%, and 0%. The EEG denes that power plants with sizes of 100 kW and above should be able to provide a remote controllability. However, as a PV plant is dened as a module and not a cluster of modules and inverters at one PCC, PV is at present not included. However, it is intended to clarify this gap in the next release of the EEG in 2012. Also, a reduction of the plant size that has to be controllable is under discussion. In terms of frequency control, all PV systems have to reduce their active power output with a droop function of 40% of their current active power output per Hertz, if the frequency exceeds 50.2 Hz. The reduced power output must remain until the frequency falls below 50.05 Hz again. All generators have to disconnect immediately, if the frequency drops below 47.5 Hz or exceeds 51.5 Hz.

Table II. Reactive power provision methods according to the German technical guideline for the connection to the mediumvoltage network. Method Fixed cos cos(P) characteristic Fixed Q Q(U) droop function Remote set values Description Fixed power factor Power factor depending on current active power output Fixed amount of reactive power Amount of reactive power depends on voltage magnitude at the local point of common coupling Set values for reactive power or power factor via remote control Response time Within 10 s Between 10 and 60 s Within 60 s

Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Distribution grid and large-scale PV deployment

M. Braun et al.

Table III. Minimum power factor according to the technical guideline for the connection to the low-voltage network. Maximum apparent power Smax Smax 3.68 kVA (ref. EN50438) 3.68 kVA < Smax 13.8 kVA Smax > 13.8 kVA
DSO, distribution system operator.

Minimum power factor 0.95 leading/lagging 0.95 leading/lagging 0.9 leading/lagging

Remark No parameter setting by DSO Parameter setting by DSO Parameter setting by DSO

The provision of reactive power is limited to times with an active power feed-in of more than 20% of the rated power of the PV system. The suggested reactive power provision methods are either a xed power factor or a cos(P) method that has to be activated within 10 s. The transmission of online set values is not yet intended. In future revisions, more advanced controls are intended to be integrated in the LV grid code.

3.2. Italian grid codes The main technical standards for the connection of electrical generators to the Italian grid are the CEI 0-16 (reference technical rules for the connection of active and passive consumers to the high-voltage (HV) and MV electrical networks of distribution companies), the CEI 11-20 (electrical energy production system and uninterruptable power systems connected to LV and MV networks), and the CEI 11-32 (electrical energy production system connected to HV network). The CEI 11-20 is mainly focused on the connection to the LV network. The technical guidelines are currently under revision, with the main aim to draw up a unique standard on HV, MV, and LV network connections within 2011. Probably, aspects regarding PV systems active and/or reactive power control regulations could be also introduced. According to new Italian feed-in tariff, PV plants going into operation from 01/2013 on must be equipped by inverters able to provide advanced network services such as voltage regulation by means of reactive power control, remote network disconnection, and voltage drop immunity. The CEI 82-25 (guide for design and installation of PV systems connected to MV and LV networks) provides the criteria for the design, the installation, and the verication of PV systems connected to the LV and MV distribution networks according to the standards CEI 0-16 and CEI 11-20. Figure 1 shows the general scheme for the connection of a production system to the LV and MV networks according to the standards CEI 11-20 and CEI 0-16, which provide also the prescriptions of interfaces and protection devices.

Figure 1. Scheme for the connection of a production system to the low-voltage and medium-voltage networks (CEI 11-20 and CEI 0-16).

Table IV. Photovoltaic power thresholds versus network voltage level. Power (kW) 100200 2003000 300010 000a
a

Network voltage level MV or LV MV MV or HV

In agreement with the distribution system operator.

LV, low voltage; MV, medium voltage; HV, high voltage.

3.2.1.1. Supply voltage characteristics. The CEI EN 50160:2010 (voltage characteristics of electricity supplied by public distribution networks) denes, describes, and species voltage quality criteria at the customers PCC in public LV, MV, and HV AC networks under normal operating conditions. Table V gives an overview about steady-state voltage criterion and permissible frequency deviations. The thresholds reported in Table V are considered about interfaces and protection devices settings for the PV systems grid connection according to the CEI standards.

3.2.1. The Italian medium-voltage grid code According to the MV grid code, the maximum power of a PV plant to be connected to an MV network depends on the grid and the local load characteristics (Table IV).

3.2.1.2. Active and reactive power control. No active and reactive power control regulations are implemented for PV systems. PV systems must not supply reactive
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

M. Braun et al.

Distribution grid and large-scale PV deployment

Table V. Medium-voltage supply characteristics (CEI EN 50160:2010). Parameter Frequency variations: for systems with synchronous connection to an inter-connected system Supply voltage variations Limit
49.550.5 Hz 47.052.0 Hz

during 99.5% of a year during 100% of the time at least 99% of the 10-min mean RMS values of the supply voltage shall be between 10% of the systems nominal voltage none of the 10-min mean RMS values of the supply voltage shall be outside the limits 15% of the declared voltage Uc

RMS, root mean square.

power, and no control features for active power are foreseen. In case of local overvoltage and overfrequency, the inverter will disconnect and reconnect once the voltage has decreased again. In general, the thresholds for power factor at delivery power point are dened in the contract for the connection of the customer to the grid in agreement with the reference technical rules (CEI 11-20 and CEI 0-16). Under specic situation, due to the typology of the customers system and/or to the grid, it can be needed to agree with the DSO a reactive power trade system different with respect to the standard one indicated by reference technical rules and authority. 3.2.2. The Italian low-voltage grid code According to the CEI 11-20, the maximum power of a PV plant that can be connected to an LV network depends on the grid and the load characteristics (compare Table VI). 3.2.2.1. Supply voltage characteristics. Table VII shows the specication for supply voltage variations at LV levels. The thresholds reported in Table VII are considered for interfaces and protection devices settings for the PV systems grid connection according to the CEI standards.

3.2.2.2. Active power control. For PV systems, no active and reactive power control regulations are implemented. PV systems must not supply reactive power, and no control features for active power are foreseen. In case of local overvoltage and overfrequency, the inverter will disconnect and reconnect once the voltage has decreased again. 3.2.2.3. Power factor. The CEI 11-20 sets the converters power factor guidelines for energy production systems and uninterruptable power systems connected to LV and MV network described in the succeeding text: 0.8 lagging power factor if the reactive power provided is in the range (20%100%) of the total installed power; leading power factor if the reactive power provided is not higher than the minimum value among 1 kvar and (0.05 + P/20) kvar, where P is the total installed power in kW; and cos = 1 in any other case (this is the requirement for grid-connected PV systems). 3.2.2.4. Prevention of islanding operation. To prevent islanding operation, the PV inverter is equipped with an islanding detection system. 3.3. United States grid codes In the USA, each utility is responsible for setting its own inter-connection requirements. With over 3000 individual utilities located in the USA, this was a challenge to the manufacturers of distributed generation (DG). In 1999, the Institute for Electrical and Electronics Engineers (IEEE) began the development of a national consensus-based

Table VI. Photovoltaic power thresholds versus distribution network voltage level. Power (kW) 6 100 100200a
a

Distribution network voltage level LV (single phase) LV LV or MV

In agreement with the DSO.

LV, low voltage; MV, medium voltage.

Table VII. Low-voltage characteristics for supply voltage variations (CEI EN 50160:2010). Parameter Frequency variations: for systems with synchronous connection to an inter-connected system Supply voltage variations Limit
49.550.5 Hz 47.052.0 Hz

during 99.5% of a year during 100% of the time 95% of the 10-min mean RMS values of the supply voltage shall be within the range of the standard nominal voltage Un 10% all the 10-min mean RMS values of the supply voltage shall be within the range of Un + 10%/15%

RMS, root mean square.

Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Distribution grid and large-scale PV deployment

M. Braun et al.

standard for inter-connecting distributed energy to electric utility distribution systems. IEEE 1547 (Standard for Interconnecting Distributed Resources with Electric Power Systems) was completed in 2003 and reafrmed without changes in 2008 [6]. This standard covers inter-connection of DG 10 MVA and under, connected at the LV or MV levels in the USA. In 2005, IEEE 1547 was included in the Energy Policy Act, which is a national law, and most utilities have used it as the basis for their inter-connection requirements. Since 2005, more than 40 states have adopted IEEE 1547 into their inter-connection rule, making it the de facto standard for inter-connection in the USA. The IEEE 1547 requirements were designed around low penetrations of DG, with the idea that the DG should disconnect from the utility system when there are any signs of trouble with utility operations such as faults, outages, and voltage sags and swells. This is to allow the utility to correct the problem before the DG would come back on line. IEEE 1547 prohibits active voltage regulation at the PCC. Because the standard only prohibits active voltage regulation, some utilities are employing a reactive power schedule or set point off of unity power factor. The standard also has relatively fast trip points for undervoltage and overvoltage and frequency excursions, yet it has no provisions for LV ride through that may be useful to enhance system stability during transmission faults. Also, IEEE 1547 only requires monitoring provisions for systems 250 kVA and greater. As the penetration of DG increases, the current requirements of IEEE 1547 may not be desirable because they could cause additional instability in the electric power system by forcing large amounts of DG to drop from the grid during minor disturbances. To address these issues and to make DG take a more active role in the operations of the power system, several efforts are underway in the USA to improve the capabilities of DG. The US Department of Energy (DOE) with help from the National Renewable Energy Laboratory and Sandia National Laboratories has developed a comprehensive approach to work with industry to address the issues that high levels of PV will bring. This approach includes development of advanced inverter technologies, updating distribution system planning practices, updating interconnection standards and codes, and documenting successful case studies on high-penetration PV projects. The IEEE has recently started a companion standard to IEEE 1547, called IEEE P1547.8 (Recommended Practice for Establishing Methods and Procedures that Provide Supplemental Support for Implementation Strategies for Expanded Use of IEEE Standard 1547) [7]. IEEE P1547.8 will provide more exibility in determining the design and processes used in expanding the implementation strategies used for inter-connecting distributed resources with electric power systems through advanced functionality such as implementation of LV ride through and voltage regulation at the distribution level. The Electric Power Research Institute and national laboratories have been working with industry in the USA

to develop a standardized set of functions for inverters [8]. The functional set includes LV ride through, curtailment, and voltage and reactive power control. In terms of technology development, the US DOE is funding several projects under the Solar Energy Grid Integration System Program to develop inverters with advanced functional and communications capability [9]. 3.4. Japanese grid codes In Japan, the technical requirements for grid inter-connection of PV systems in LV distribution systems are regulated by the interpretation of technical standards for electrical equipment and the guideline of power quality for grid inter-connection. 3.4.1. Voltage regulation According to the Electric Utilities Industry Law, voltage at inter-connection point of PV system must be within 101 6 V. 3.4.2. Power factor Power factor of PV systems must be more than 0.85 lagging, and leading power factor is prohibited to prevent the voltage rise. 3.4.3. Operation at momentary voltage drop At the momentary voltage drop, PV systems have to continue generating power or return quickly after the voltage drop. 3.4.4. Power control to reduce voltage rise When the voltage at inter-connection point of PV system exceeds the upper limit, the PV system is required to control reactive power output or to reduce active power output. A typical control procedure is as follows: PV systems provide reactive power to reduce the voltage rise with keeping the power factor more than 0.85. If the reactive power control is not enough to regulate the voltage, PV systems reduce active power output with maintaining the power factor 0.85. 3.4.5. Prevention of islanding operation To prevent islanding operation, PV systems must have both active and passive islanding detection systems. PV systems have to disconnect if the frequency drops 3% below the nominal frequency or exceeds +2% of the nominal frequency. 3.5. Belgian grid codes In Belgium, grid codes for the connection of PV to the distribution system are issued by Synergrid, the Belgian federation of grid operators for electricity and natural gas. The technical requirements are bundled in the document C10/11 [10]. The code largely refer to existing standards and guidelines, in particular, the European EN 50438 [11] for the connection of microgenerators in parallel with public LV distribution networks. The C10/11 is applicable for all distributed generators connected to the distribution grid.
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

M. Braun et al.

Distribution grid and large-scale PV deployment

Single-phase installations are allowed until 20 A (approximately 5 kVA). For larger installation, the injected power needs to be distributed over the phases, and the overall phase imbalance must not exceed 20 A. For example, a 10-kVA installation of two 5-kVA inverters connected to two different phases would be allowed, but a 6-kVA installation of one 6-kVA inverter would not be allowed. For prevention from unintended islanding, C10/11 makes a difference between installations of 10 kVA and those >10 kVA. For installations of 10 kVA, an automatic disconnection device is allowed. For this device, C10/11 explicitly requires compliance with the German DIN V VDE V 0126-1-1 from February 2006 [12]. For larger installations, a disconnect switch accessible to the grid operator is required. The current requirements according to C10/11 do not foresee any contribution to a more efcient grid management from the PV inverter. No control features for active power are foreseen. For PV installations >1 MVA, the DSO can request the injection of reactive power up to 33% of rated power. In case of local overvoltage, the inverter will disconnect and reconnect once the voltage has decreased again. With increasing PV penetration, grid operators report that, already today, PV installations regularly disconnect because of overvoltage. 3.6. Australian grid codes Over 90% of electricity consumers and consumption in Australia falls within the jurisdiction of the Australian National Electricity Market, which covers all states and territories other than Western Australia and the Northern Territory. The National Electricity Law and Rules governing the National Electricity Market provide high-level guidance on embedded generation (EG) connection to the distribution network. An EG is dened in the rules as being a generator connected to a distribution network, which does not have direct access to a transmission network. The National Electricity Rules lay out detailed obligations on network service providers for providing connection to EGs, and associated compensation and charging principles [13]. Generators of less than 5 MW are exempt from market registration, dispatch processes, and market settlements. State jurisdictional arrangements can also include specic access requirements and codes. Relevant standards include AS/NZS 5033 Installation of photovoltaic (PV) arrays and AS4777 Grid connections of energy systems via inverters, which species the requirements for inverters, with ratings up to 10 kVA for single-phase units or up to 30 kVA for three-phase units, as well as requirements for grid associated protection devices, for the injection of electric power through an electrical installation to the electricity distribution network. Many of the specic requirements, however, fall within the jurisdiction of the Distribution Network Service Providers. In particular, the connection of a grid-connected inverter system is subject to formal agreement with the electricity distributor. Their requirements can vary
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

signicantly although there are generally many similarities. State jurisdictional efforts may also be signicant. For example, ActewAGL services Australias capital city, Canberra, and favorable policy support for PV and customer demographics have combined to see considerable penetrations of domestic PV systems. ActewAGL species a maximum rating of an inverter allowed for single-phase installations of 10 kVA. PV installations with greater capacities must be three-phase systems. The output voltage of the PV system must fall within the range specied in ActewAGLs Service: 226254 V. The power factor of the PV system shall not exceed the limits dened in AS4777, and all protection equipment associated with a grid-connected PV system must be designed, installed, and tested to ensure that islanding does not occur. Photovoltaic installations must disconnect from ActewAGLs distribution network in the event that one or more phases of the distribution network is lost. Finally, ActewAGL limits the capacity of large PV installations connected without a dedicated transformer, such that the combined maximum capacity of all PV installations connected without a dedicated transformer must remain less than 30% of the rated capacity of the transformer they are connected to [14]. Many aspects of the grid connection framework are now under review for reasons that include the increasing application of EG, particularly with PV systems. Options being pursued include standardized standing offers for connection of small embedded generators [13], and changes to AS4777 to provide more exible inverter options to address high penetration issues. 3.7. Comparison of different grid codes This review of grid codes clearly shows the differences between countries. Italy, Belgium, and Australia do not have any active or reactive power control regulations. In the USA, a new standard IEEE P1547.8 is in preparation that will provide more exibility for interconnecting distributed resources with electric power systems through advanced functionality such as implementation of voltage regulation at the distribution level. Furthermore, utilities are testing different approaches for improved integration of PV. In Japan, PV systems have to follow a voltage regulation scheme in LV systems that is activated as soon as the upper limit 101 + 6 V is reached. Firstly, reactive power is activated up to a power factor of 0.85 lagging. Secondly, if the voltage is not yet kept within limits, active power of the PV system is reduced. In Germany, the present grid code at MV requires most functionality from PV systems. This includes active power control depending on the locally measured frequency for positive frequency support as well as active power control by set values from the DSO in case of security risks. In addition, various reactive power control settings can be requested by the DSO: central control by DSO set values, local static settings for reactive power or power factor, and

Distribution grid and large-scale PV deployment

M. Braun et al.

local functions for reactive power depending on PV active power generation and terminal voltage measurements. At present, the LV grid code does not require any of these control features. However, a new LV grid code is intended to be released in 2011 providing similar control functionalities as at MV. In all countries, new releases are in preparation to deal with the impact of PV to the distribution system properly by using control functionalities of PV inverters. However, these efforts are mainly country-specic. Most important is a harmonization over different countries. This has already been initiated in Europe by DERlab and is now being pushed forward by the European Network of Transmission System Operators (ENTSO-E; see next section). Leading research institutes from 11 European states have founded in 2008 the association DERlab e.V. as an independent laboratory for the grid integration of distributed power generation. DERlab develops joint requirements and quality criteria for the connection and operation of distributed energy resources (DER) and is strongly supporting the consistent development of DER technologies. DERlab offers testing and consulting services for DG to support the transition towards more decentralized power systems [15]. The DERlab Data Base of European DER Inter-connection Specications aims at creating more transparency in the clutter of the many different specications in order to support European harmonization. DERlab also started initiatives to develop a common European pre-standard on DER interconnection [16,17].

Currently, the largest share of installed PV capacity in Europe is of type A or type B. Some of the requirements are feasible and can be easily implemented into the PV inverter. Others will have a high cost impact. The most important requirements for PV installations of types A and B are the requirement of reactive power control capabilities and the capability to operate in frequency sensitive mode (power control). Notably, ENTSO-E does not make any statement on the actual provision of these services but only requires the capability to provide them, in case they may be required at some point in the future. The question, if and how these ancillary services will be remunerated, is not treated in the network code. Therefore, integrating them with the PV system per default, that is, without knowing if and when these services may be called, would increase the costs of PV systems. Finally, the draft network code is not clear in how far existing generators will have to comply with the new requirements. Because of the large number of existing installations, the retrots required for this purpose seem hardly possible in practice.

4. SELECTED CASE STUDIES FOR AN IMPROVED GRID INTEGRATION OF PHOTOVOLTAICS


This section describes how PV can be integrated into the distribution grid. The present approach is normally grid reinforcement at high penetration levels. However, the use of smart inverter control capabilities allows reducing grid reinforcement costs, for example, by using reactive power or reducing active power to keep voltage limits. 4.1. Case study for a German low-voltage grid The following case study has already been published in [20]. This section provides a summary of the assumptions, method, and results. 4.1.1. Simulation assumptions An LV system section with a total of 54 residential customers and a high local PV penetration were analyzed with PowerFactoryW from DIgSILENT. Each of the residential costumers was equipped with a scalable PV system. The investigated LV section with its weakest point PCC is depicted in Figure 2. It only provides a single line representation of the LV grid without giving each connection point. The simulations were performed for a period of one week in 15-min resolution. For all 54 scalable PV systems, a time series with real measured solar irradiance data for the location Kassel was used. The load proles used for the residential homes are statistically derived from a real measured load prole [21]. The MV level was neglected.
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

3.8. European pilot grid code In 2010, the ENTSO-E [18] issued the draft version of a new grid code, with the latest update published in March 2011 [19]. Once nalized and approved, this network code will rule the requirements for grid connection applicable to all generators for transmission and distribution grids at all voltage levels. Generators are subdivided into the three categories synchronous generating units, power park modules, and offshore generators, where PV systems belong to power park modules. Within this category, distinction is made between four different types, depending on power and voltage level. The type denitions differ for the different synchronous zones in Europe. The thresholds for type classication (nominal power P and voltage U) in Continental Europe are as follows: type type type type A: P 400 W B: P 0.1 MW C: P 10 MW D: P 10 MW and U 110 kV

For the other synchronous areas, partly different thresholds have been proposed; however, these values may still be adapted during the review and consultation process. Therefore, we refer to the latest version on the web site of ENTSO-E (www.entsoe.eu).

M. Braun et al.

Distribution grid and large-scale PV deployment

The following scenarios were investigated: Scenario 1 The PV systems always feed-in their maximum active power. There is no provision of reactive power. Scenario 2.1 The PV systems provide reactive power according to the technical guideline for the connection to the LV network. A xed power factor method was used with a minimum power factor of 0.95. Scenario 2.2 The PV systems provide reactive power according to the technical guideline for the connection to the LV network. A cos(P) method with a minimum power factor of 0.95 was used. Scenario 3 Reactive power will only be provided if the local PCC voltage magnitude exceeds a threshold value of +3% of the nominal system voltage. In this case, the reactive power output of the inverter will be

adjusted to maintain the local grid voltage. If the reactive power output is not sufcient to maintain the local grid voltage, a reduction of the active power output is also possible. 4.1.2. Simulation method The installed PV capacity at each PCC was up-scaled simultaneously until either the +3% voltage threshold was exceeded at any PCC within the grid or local equipment overloading occurred. The overloading limitation was set to 130% of its rated power. By applying the different scenarios, we could determine the effect of the different reactive power provision methods to the hosting capacity of the grid. Figure 3 compares the bandwidth of the active power output, the reactive power output, and the local voltage at the grids weakest PCC. In this case, an installed PV capacity of 5 kVA was assumed for all scenarios.

Figure 2. Investigated low-voltage section. Weakest point of common coupling (PCC) is marked. All other 53 PCCs are not depicted.

Figure 3. Bandwidth of the active and reactive power output of the inverter and the local voltage magnitude over one week at the grids weakest point of common coupling (PCC).
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Distribution grid and large-scale PV deployment

M. Braun et al.

Figure 4 compares the maximum hosting capacity of the grid for PV capacity. It is shown that with no additional reactive power provision the hosting capacity of the grid was 3.5 kWp/PCC. Using the capability of additional reactive power consumption, according to the technical guideline for the connection to the LV network, we can increase the hosting capacity of the grid to 5.5 kWp. An even higher hosting capacity could be achieved with the adaptive reactive power provision of scenario 3 (7.5 kWp/PCC). Because of its decentralized voltage control functionality, this method allows to increase the local PV capacity until the loadings limitations of the network equipment are reached. This signicant increase of the systems hosting capacity has to be beard by occasional active power feed-in losses. 4.2. Case study for an Italian distribution system The Italian National Agency for New Technologies, Energy and Sustainable Economic Development (ENEA) in cooperation with University of Palermo performed a technical and economical feasibility study as well as a preliminary design of a demonstrator of a distribution electrical system, on the basis of renewable energies for the transition towards active networks. The study wants to contribute to the process of identication of new reference modes and to the standardization of the relevant solutions aiming at the penetration of DG over the networks. 4.2.1. Simulation assumptions The study is carried out on a small section of a real MV/LV distribution system of the research center ENEA of Casaccia [22,23] (Figure 5). Substation 20 is equipped with two MV/LV transformers with a rated voltage of 630 kVA. Each of them is connected on its MV side through a delta connection and on the LV side through a wye connection with neutral conductor. Substation 16 is equipped with an MV/LV transformer with rated power of 1000 kVA having the same typologies of connections as

Substation

Figure 5. Simplied schema of the network in the area La Capanna. The main medium-voltage and low-voltage lines are indicated.

substation 20, and it exclusively supplies, through an LV feeder, an experimental concentrated solar power plant. Substations 16 and 20 are supplied through two radial main MV lines, coming from the HV/MV substation and named 1E and 2E. The supply lines of the electrical switchboards located in the buildings come from substation 20. From the same substation comes out an LV line supplying the substation 15. Into the latter substation, a further LV switchboard is installed. It supplies the terminal switchboards of the buildings. Within the building where substation 20 is located, also two compensation switchboards are installed. Each of them is controllable in the following steps: 51020 kvar. The yearly consumption of electrical energy is about 1.38 GWh, with a peak of 167 MWh in August and a minimum of 73.5 MWh in December. The yearly reactive power consumption is about 0.7 Gvarh, with a maximum of 111 Mvarh in July and a minimum of 27 Mvarh in December. Not only considering the availability of energy resources that can be fruitfully exploited but also taking into account all the other main aspects (constraints, availability of spaces, impact issues, territorial and functional specicity, etc.) that the real implementation requires, among the different renewable sources, it is considered to employ the solar resource through PV conversion, as well as the wind source and the cogeneration through the use of gas supplied microturbines. Two scenarios, which are characterized by different coverage of the yearly electrical needs through renewable sources, were considered. The simulations were carried out for the following scenario (coverage of about 34% of the yearly electrical load) by means of the installation of the following: PV plants for an overall power of 45 kWp for yearly production of about 68 kWh; one wind microgenerator of 20 kW; and one microturbine for the combined production of electrical and thermal energy, which is able to deliver 50 kW of electrical power (and a thermal energy of 100 kW).
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Figure 4. Maximum hosting capacity of the investigated lowvoltage network for photovoltaic capacity.

M. Braun et al.

Distribution grid and large-scale PV deployment

4.2.2. Simulation method A model of the network has been implemented using NEPLAN simulation software, starting from the supply HV/MV point up to the LV loads represented by the switchboards of the buildings. The model, validated through a comparison with the data coming from the output data of the initial or reference scenario (passive system), has been used for the load ow analysis (in particular working conditions) and of short circuit (for three-phase faults in different nodes of the system) both for the initial scenario (passive), and for the previously indicated scenario of active network (with DG), in the aim of evaluating the systems performance. For the two hypothesized working conditions (maximum loading of 702 kW and minimum loading of 180 kW), the results of the load ow simulations are shown in Tables VIII and IX for the active/reactive power Pload/ Qload of the load, the active/reactive power PgRES/QgRES of the generated renewable energy from DG, the active/reactive power Pggrid/Qggrid from the superior grid, and the active power losses PTOT. Figures 6 and 7 show the voltage proles in the different network nodes, sorted in monotonically decreasing order, both for the condition at maximum load and for the minimum load condition. The analysis of the results have shown that a reconguration of the grid (passive versus active) is needed to allow a large penetration of DG assuring the reliability and the quality of the electric service; the connection of new generators (active network) produces a reduction in losses (both active and reactive power), which is obviously more sensitive in the condition of maximum load; and the voltage prole is more regular in both operation conditions. A similar analysis for the transition of real distributed system towards an active network [24,25] is carried out on the isolated MV grid of the Island of Pantelleria located in the Mediterranean area close to Sicily.

4.3. United States case studies In the USA, the integration of high penetrations of PV is just beginning. Typically, utilities have allowed integration of penetrations up to 15% without the need for detailed inter-connection studies, where PV penetration is dened as the aggregate ratings of the PV systems on a circuit (or line segment) divided by the peak load of the circuit (or line segment). Utilities use 15% circuit (or line segment) penetration as a conservative level meant to limit the impact of the PV system on voltage regulation, equipment ratings, grounding, protection coordination, and risk of unintentional islanding. We show in the succeeding text several case studies of PV projects in the USA that are much greater than 15% penetration. These systems are being

Figure 6. Condition at maximum load: voltage proles for the two congurations (passive system and active one).

Figure 7. Condition at minimum load: voltage proles for the two congurations (passive system and active one).

Table VIII. Maximum load condition: load ow results. Cong. Passive system Active system Pload (kW) 702 702 Qload (kvar) 716 716 Pggrid (kW) 709 611 Qggrid (kvar) 647 616 PgRES (kW) 98 QgRES (kvar) 27 PTOT (kW) 7 6

Table IX. Minimum load condition: load ow results. Cong. Passive System Active System Pload (kW) 180 180 Qload (kvar) 69 69 Pggrid (kW) 180 124 Qggrid (kvar) 60 34 PgRES (kW) 56 QgRES (kvar) 27 PTOT (kW) <1 <1

Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Distribution grid and large-scale PV deployment

M. Braun et al.

studied to identify issues and develop solutions to high penetrations of PV integrated with electrical distribution systems. 4.3.1. Big Island, Hawaii On the Big Island, Hawaii, PV installations have been increasing at a signicant rate. The Hawaiian Electric and Light Company (HELCO) operates the power system and has been working to address higher levels of distributed PV integrating into their system. HELCO has several distribution circuits that operate at over 50% PV penetration. Because the Big Island is a relatively small grid, it has larger uctuations in frequency than experienced in the continental USA under normal operations. HELCO is now requiring that all distributed PV including systems under 30 kW be installed with adjustable voltage and frequency trip points so that they will be coordinated with HELCOs underfrequency load shedding schemes (private discussions with HELCO, January 2011). 4.3.2. Kauai, Hawaii The power system on the Hawaiian island of Kauai is operated by the Kauai Island Utility Cooperative (KIUC). KIUC has recently installed a 1-MW PV system on a distribution circuit that represents a 100% penetration level. During sunny days, this system consistently backfeeds power to the substation transformer (private discussions with KIUC, January 2011). 4.3.3. San Antonio, Texas CPS Energy is the electric utility in San Antonio, Texas. CPS has recently inter-connected a 14-MW PV plant onto two adjacent distribution circuits within their service territory, which currently have a penetration level of approximately 80% on each circuit (private discussions with CPS Energy, January 2011). 4.3.4. Fort Collins, Colorado In Fort Collins, Colorado, a 5.2-MW PV system is installed on the Colorado State University Foothills Campus and inter-connects onto the local utilitys (Xcel Energy) distribution system. The feeders average daily peak load was 7.5 MVA in 2010, with a peak demand load of 9.1 MVA and a minimum daytime load of 3.1 MVA, which equates to approximately 47% PV penetration. The utility conducted an inter-connection study to determine possible impacts of the system. The results indicated there were no major concerns of the PV systems other than possible HV during light loads. If HV ever becomes a problem, the utility will approach the issue by rst adjusting the voltage regulators on the line to stabilize the voltage levels. If that does not solve the issue, the inverters will be changed to absorb reactive power at xed levels to reduce the feeder voltage. If that does not reduce the voltage sufciently, the inverters will be disconnected from the system [26].

4.3.5. Fontana and Porterville, California Southern California Edison (SCE) is the utility that operates a majority of the electrical power system in the areas that surround the city of Los Angeles, California. SCE is currently in the process of inter-connecting 500 MW of distributed PV systems in their service territory. In Fontana, California, SCE has installed a 2.3-MW PV system on a commercial warehouse roof. This system represents a 23% PV penetration. Because there are a large number of warehouses in the area surrounding this system, it is expected that several additional megawatts of PV systems may be installed on the same circuit, potentially increasing penetration to nearly 63% on that circuit. In Porterville, California, SCE has installed a 5-MW PV system on a long rural circuit that represents a penetration level of over 100%. These systems are being evaluated as part of a research project sponsored by the US DOE, and California Public Utility Commission, and will be retrotted with inverters that have the ability to control reactive power and curtail PV power output to regulate voltage. It is expected that these inverter upgrades will allow additional PV systems to be integrated into the distribution circuits [27]. 4.4. Case study for a Japanese distribution grid To maintain the local supply voltage, power control of PV systems is required according to Japanese grid-interconnection code. However, distribution line voltage is usually different at each location so that large imbalance of output active power between PV systems may occur. The control method to share reactive power has been studied in [28,29]. 4.4.1. Autonomous control method The problem of the present method is that the number of PV systems producing reactive power is limited. Therefore, autonomous control method of PV systems to share reactive power between them without information and communication technology (ICT) was proposed. In the proposed method, a new voltage threshold (e.g. 106.5 V), which is less than the present voltage upper limit 107 V, is introduced. Figure 8 shows an overview of the proposed method to reduce the voltage rise. Once the voltage at the PV system reaches the new voltage threshold, the PV system begins to inject reactive power up to a limit of power factor (e.g. 0.85). Then, it is expected that more numbers of PV systems contribute to restrain the voltage rise as shown in Figure 9. As a result, imbalance of output active power between PV systems is improved. By the simulation analyses using Japanese typical distribution line model of residential area, it was conrmed that improvement rate of PV power generation energy per day reaches 7% at maximum compared with present method [28]. 4.4.2. Remote control method To enhance applicability, a remote control method with ICT infrastructures was proposed [29]. The proposed method was designed taking account of controlling reactive power
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

M. Braun et al.

Distribution grid and large-scale PV deployment

Figure 8. Proposed method for restraining voltage rise in distribution line.

Figure 10. Procedure of remote control method (start of voltage regulation).

Figure 9. Comparison of photovoltaic (PV) power output between present method and proposed method.

between unspecied PV systems with the minimum frequency of the communication. Figure 10 shows the owchart of the proposed method. If there is a PV system that should reduce active power to suppress the voltage rise (DG1), it requests another PV system (DG2) to share reactive power by using ICT infrastructures. When the active power of DG1 is recovered by the reactive power injection from DG2, DG1 requests DG2 to x their reactive power output to minimize the reactive power output of DG2. Demonstration tests of the proposed remote control method are carried out with an actual scale experimental distribution system at CRIEPI Akagi Testing Center. Figure 11 shows the conguration of the experimental distribution system and ICT infrastructures. DG systems are installed at four different sites and connected to 6.6-kV experimental distribution line through a three-phase 200-V LV distribution line. Each DG system is a 20-kW invertertype power generator. Each DG system is also connected to
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Figure 11. Experimental distribution line and information and communication technology infrastructures. DG, distributed generation; SDI, Supply and Demand Interface.

the communication network through a communication device Supply and Demand Interface. Figure 12 shows the comparison of DG systems output power between the proposed remote control method and the conventional local autonomous control method when each DG system operates with 15 kW. From the results, it is clear that active power output of DG3 is improved remarkably by sharing reactive power with DG4. 4.5. Belgian case study With the recent strong growth of PV in Belgium, distributed PV systems regularly experience disconnection

Distribution grid and large-scale PV deployment

M. Braun et al.

Figure 12. Comparison of distributed generation (DG) output power during voltage control. Figure 13. Frequency distribution for several critical mediumvoltage and low-voltage cabinets within the range of 402 and 420 V (nominal 400 V) [34].

due to overvoltage, and in several cases, expensive grid reinforcement is required in order to avoid congestion of cables or transformers. A recent initiative for demonstrating and testing of how PV can actively contribute to a more intelligent grid management is the MetaPV project. MetaPV is a demonstration activity funded by the European Commissions Seventh Framework Programme (grant agreement TREN/ FP7EN/239511/METAPV). It was launched by 3E jointly with the Belgian DSO Infrax and in cooperation with the AIT, SMA Solar Technology, the University of Ljubljana, and LRM. Within MetaPV, active grid support from concentrated PV will be demonstrated in two contexts: in residential/ urban areas with 512 kW of PV with 4 kW on average and in industrial areas with 6.2 MW of PV with systems around 100 to 1000 kW. In addition to these systems with active grid support functions, several megawatts of classical PV are already operational in the demonstration areas. The control capacities to be implemented into PV inverters and demonstrated are active voltage control, fault ride-through capability, autonomous grid operation, and interaction of distribution system control with PV systems. The MetaPV project runs from October 2009 until March 2014. Currently, MetaPV is in the preparation phase for the large-scale demonstration. Grid and inverter simulations are being carried out and a measurement campaign on the selected grids has been launched, rst, to monitor the status quo and, later, the situation with inverters for grid support being operational. Some rst results are available from the preparatory phase: 110 power measurement devices have been installed for monitoring the electrical parameters in the MV to LV cabinets in the demonstration area. Figure 13 shows results from one week in summer 2010 (week 35) for four MV to LV cabinets. The measurements show that voltages are regularly very high, already today, illustrating the necessity and also urgency of grid support from PV to avoid costly grid reinforcements. The most important benets that PV can bring to the distribution grid have been identied. The most relevant

features are reactive power, power factor control, active power control, dynamic grid support, and in combination with a storage unit, asymmetry mitigation. In June 2011, 125 small PV systems with inverters for grid support are installed with a capacity of 506 kW, distributed over 81 houses in the demonstration areas Lommel and Opglabbeek. Figure 14 shows an example of a voltage drop diagram in the case study grid. It is clearly visible how PV and wind penetration levels compensate the voltage drop from loads and even increase the voltage up to given limits of 110% of the nominal voltage. The main work to be done in the coming years covers the following topics: Interaction with the central dispatch: How should local voltage control by the inverter and supervisory grid control interact?

Figure 14. Example of voltage drop diagram for the worst-case low load and high generation (voltage set point 1.05 p.u. at the substation): year 2010 [35].
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

M. Braun et al.

Distribution grid and large-scale PV deployment

Distribution of storage systems: Storage systems may allow reducing overvoltages and congestion and mitigating phase imbalance; however, no common rules for the placement of storage exist in distribution system management. Quantifying economic value: The added value that active PV may bring to grid operations has often been quoted, but very little quantitative information is available. Operation/monitoring/demo: In line with the project objectives, the MetaPV consortium wants to demonstrate active grid support from PV on a large-scale and in a real-world environment. The test installations and interaction with the grid operations are planned to become fully operational in the course of 2012. 4.6. Australian case study The Australian PV market is less developed than those of other countries. However, there has been very rapid growth in the deployment of domestic PV systems over the last several years under a range of State and Federal policy initiatives. It is estimated that around 350 MW was installed in 2010, up from around 75 MW in 2009, and that over 7% of owner-occupied houses in the largest Australian State, NSW, now have a rooftop PV system [30]. Different jurisdictional policy support measures, a number of government programs such as Solar Cities, and installer market strategies have seen particular concentrations of distribution connected PV in some Australian regions. Almost all of this PV has been installed in domestic applications with system sizes of less than 5 kW, although there are a growing number of larger systems now being installed. One example of signicant and growing PV penetrations is the remote town of Alice Springs in the Australian Northern Territory, which now has over 3 MW of PV on a system with a peak load of less than 60 MW. This is an isolated grid supplied from gas-red generation. The NT Power and Water Authority, which is responsible for the electricity system, has noted some emerging PV network issues and has taken a number of management actions. For example, the Northern Territorys Power and Water Authority now required PV inverters connected to the Alice Springs electricity network to be congured such that they stay on line for frequency excursions of 50 4 Hz because of issues arising from the loss of PV systems at times of low-frequency excursions [31].

should behave as passive as possible. In contrast, others demand an active participation in grid control. The review of current grid codes in some of the leading PV countries shows a diverse approach and a signicant need for harmonization. A number of country-specic case studies show different approaches for improved integration of PV systems in the distribution grid. A recent study [32] on the distribution grid extension costs in Germany due to PV and wind integration (scenario 2020: 52-GW PV and 36-GW wind) presents gures in the range of 2127 billion Euros. This is caused by 240 000-km lines at LV substation and 140 000-km lines at MV substation as well as 33 GVA at LV/MV substations and 30 GVA at MV/HV substations. These are calculations based on conventional grid planning procedures. As presented in the German case study, smart PV inverter functionalities using active and reactive power control enable reducing grid reinforcement costs signicantly. As the present PV market share leader, Germany already has invested billions of Euros in distribution grid extensions and reinforcements due to PV installations. Grid codes are being developed to allow active and reactive power control that can reduce these costs signicantly. However, more than 17-GW PV are already installed. More than 90% do not have any of these power control capabilities. Other countries that will see future increase of PV in distribution grids should learn from these experiences and request smart PV inverters with active and reactive power control capabilities. In addition, they should consider these controls in grid planning procedures. On a global scale, this allows signicant integration cost savings. In the International Energy Agency Photovoltaic Power System Task 14, recommendations will be elaborated that will provide guidelines for grid code developments in different countries as well as for harmonization efforts [33].

REFERENCES
1. European Photovoltaic Industry Association (EPIA). 2015 Global Market Outlook for Photovoltaics until 2015, Brussels, 2011. 2. Renewable Energies Agency. EEG Stammdaten, Berlin, 2010. 3. BDEW. Technische Richtlinie Erzeugungsanlagen am Mittelspannungsnetz, Berlin, June 2008. 4. VDE FNN. Technical standard for the connection and parallel operation of generators at low voltage levels, VDE-AR-N 4104. Frankfurt, August 2011. 5. VDE FNN. Rahmenbedingungen fr eine bergangsregelung zur Frequenzabhngigen Wirkleistungssteuerung von PV-Anlagen am NS-Netz. Berlin, March 2011. 6. IEEE 1547-2008. Standard for Interconnecting Distributed Resources with Electric Power Systems. 7. IEEE 1547.8. Draft Recommended Practice for Establishing Methods and Procedures that Provide

5. CONCLUSIONS
The installed capacity of PV systems has increased at a much faster rate than what the development of adequate grid codes was able to cope with. This has caused totally different grid connection requirements in different countries. Some still have an approach that PV systems
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Distribution grid and large-scale PV deployment

M. Braun et al.

8.

9.

10.

11.

12.

13.

14. 15. 16.

17.

18.

19.

20.

21.

Supplemental Support for Implementation Strategies for Expanded Use of IEEE Standard 1547. Seal B. Smart Inverter Communications Initiative. 4th International Conference on Integration of Renewable and Distributed Energy Resources, Albuquerque, NM, December 2010. Bower W. Solar Energy Grid integration Systems (SEGIS): Goals, Processes and Impacts. 4th International Conference on Integration of Renewable and Distributed Energy Resources, Albuquerque, NM, December 2010. Specieke Technische Aansluitingsvoorschriften voor Gedecentraliseerde Productie-installaties die in Parallel Werken met het Distributienet. Synergrid, C10/11, Rev. 12, May 2009. EN 50438. Requirements for the Connection of Microgenerators in Parallel with Public Low-voltage Distribution Networks. CENELEC, December 2007. DIN V VDE V 0126-1-1. Selbstttige Schaltstelle Zwischen einer Netzparallelen Eigenerzeugungsanlage und dem ffentlichen Niederspannungsnetz. VDE: Berlin, Germany, February 2006. Energy Networks Association (ENA). Embedded Generation: ENA Policy Framework Discussion Paper, Barton, Australia, November 2008. Actew AGL. Guidelines for Grid-connected Photovoltaic Installations via Inverter. Released 30 December 2009. DERlab e.V. http://derlab.eu/media/pdf/docs/D2.1_ Rev7f_PU.pdf [accessed on 11 May 2011]. Brndlinger R, Bletterie B. Key Needs, Priorities and Framework for the Development of a Common European Pre-standard on DER Interconnection. DERlab. Deliverable 2.1, 2008. Strauss P (ed.). DER Inverter White Book. DERlab. Deliverable 2.8, Kassel, Germany, 2009. http://www. derlab.eu/media/pdf/docs/DERlab_D2.8_whitebook_ static_converters_rev0.pdf [accessed on 12 May 2011] European Network of Transmission System Operators for Electricity (ENTSO-E). Draft Requirements for Grid Connection Applicable to all Generators, Brussels, March 2011. Brussels, Belgium, http://www.entsoe.eu [accessed on 12 May 2011] Draft Requirements for Grid Connection Applicable to all Generators, ENTSO-E, 22 March 2011. http:// www.entsoe.eu/ [accessed on 12 May 2011]. Stetz T, Yan W, Braun M. Voltage Control in Distribution Systems with High Level of PV-penetration Improving Absorption Capacity for PV Systems by Reactive Power Supply. 25th European Photovoltaic Solar Energy Conference, Valencia, 2010. Braun M, Bdenbender K, Thomas U, Schmiegel A, Schuh H, Magnor D, Sauer DU, Marcel JC. Improving PV-integration into the Distribution

22.

23.

24.

25.

26.

27.

28.

29.

30. 31.

32.

GridContribution of Multifunctional PV-battery systems to Stabilised System Operation. 25th European Photovoltaic Solar Energy Conference and Exhibition, Valencia, 2010. Favuzza S, Graditi G, Ippolito MG, Massaro F, Musca R, Riva Sanseverino E, Zizzo G. Transition of a Distribution System towards an Active Network. Part I: Preliminary Design and Scenario Perspectives. ICCEP: Ischia (Italy), 1416 June 2011. Cosentino V, Graditi G, Favuzza S, Ippolito MG, Massaro F, Riva Sanseverino E, Zizzo G. Transition of a Distribution System towards an Active Network. Part II: Economical Analysis of Selected Scenario. ICCEP: Ischia (Italy), 1416 June 2011. Favuzza S, Graditi G, Ippolito MG, Massaro F, Musca R, Riva Sanseverino E, Zizzo G. From Fuel Based Generation to Smart Renewable Generation: Preliminary Design for an Islanded System. Part I: Technical Issue and Future Scenarios. CIRED: Frankfurt (Germany), 69 June 2011. Cosentino V, Graditi G, Favuzza S, Ippolito MG, Massaro F, Riva Sanseverino E, Zizzo G. From Fuel Based Generation to Smart Renewable Generation: Preliminary Design for an Islanded System. Part II: Selection of Future Scenario and Economical Issues. CIRED: Frankfurt (Germany), 69 June 2011. Coddington M, Baca D, Kroposki B, Basso T. Deploying High Penetration Photovoltaic SystemsA Case Study. IEEE Photovoltaic Specialist Conference, Seattle, WA, June 2011. Mather B. Analysis of High-penetration Levels of PV into the Distribution Grid in California. High Penetration Solar Forum, San Diego, CA, March 2011. Hirabaru Y, Hatta H, Kobayashi H. Development of a Distribution System Voltage Control Method for PV SystemsA New Reactive Power Control Method for Restraining Voltage Rise. Proceedings of the 17th International Photovoltaic Science and Engineering Conference, 2007. Kobayashi H, Hatta H. Reactive Power Control Method between DG Using ICT for Proper Voltage Control of Utility Distribution System. Proceedings of the IEEE PES General Meeting, 2011. Morris N. State of the Australian PV Market. Presentation to ATRAA2011, Melbourne, May 2011. Centre for Energy and Environmental Markets (CEEM). A Case Study of High Penetration of PV Systems in Electricity Grids: The Alice Springs Electricity Supply System. Report for the Australian PV Association, May 2011. E-Bridge Consulting. Abschtzung des Ausbaubedarfs in deutschen Verteilungsnetzen aufgrund von Photovoltaikund Windeinspeisungen bis 2020. Gutachten im
Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

M. Braun et al.

Distribution grid and large-scale PV deployment

Auftrag des BDEW Bundesverband der Energie- und Wasserwirtschaft e.V. 16 March 2011. 33. Mayr C, Brndlinger R, Fechner H, Braun M, Ogimoto K, Frederiksen K, Kroposki B, Graditi G, MacGill IF, Turcotte D, Perret L. Bringing Together International Research on High Penetration PV in Electricity Grids The New Task 14 of the IEA-PVPS Programme. 4th International Conference on Integration of Renewable and Distributed Energy Resources, Albuquerque, USA, 2010.

34. Vandewaerde D, Truyens F, Lemmens E, Deprez W, Woyte A, Dexters A. Technical Site Description of the Demonstration Zone Including Results from Measurements. MetaPV, Deliverable 1.1, December 2010. 35. Bletterie B, Gorek A, Uljani B, Blai B, Woyte A, Vu Van T, Truyens F, Jahn J. Enhancement of the Network Hosting CapacityClearing Space for/with PV. 25th European Photovoltaic Solar Energy Conference and Exhibition, Valencia, 2010.

Prog. Photovolt: Res. Appl. (2011) 2011 John Wiley & Sons, Ltd. DOI: 10.1002/pip

Das könnte Ihnen auch gefallen