3

Photovoltaic module stability and reliability

D. Jordan; S. Kurtz    National Renewable Energy Laboratory (NREL), Golden, CO, United States

Abstract

Profits realized from investment in photovoltaic will benefit from decades of reliable operation. Service life prediction through accelerated tests is only possible if indoor tests duplicate power loss and failure modes observed in fielded systems. Therefore, detailing and quantifying power loss and failure modes is imperative. In the first section, we examine recent trends in degradation rates, the gradual power loss observed for different technologies, climates and other significant factors. In the second section, we provide a summary of the most commonly observed failure modes in fielded systems.

Keywords

Reliability; Durability; Degradation rates; Failures; Field test; Aging

Acknowledgement

This work was completed under contract no. DE-AC36-99GO10337 with the US Department of Energy.

3.1 Introduction

As the installed capacity of the photovoltaic (PV) industry is approaching 200 GW, investors seek confidence in long-term PV performance. Profitable investment in PV will benefit from decades of reliable operation. The discipline of reliability has been well established and has led to significant quality increases for many types of products, including PV systems; however, the application to PV modules is challenging for several reasons. The desired long lifetime prevents direct verification, especially because of the short product development life cycle. In addition, the size and cost of PV modules typically inhibits testing of a large number of samples. Furthermore, PV modules are complex and can display a variety of failure modes. Because of the synergistic nature of the different failure modes an understanding of the underlying physical and chemical processes is essential. Finally, the use environment can be highly variable, leading to a climate dependency of different failure modes. As the industry enhances service life prediction, improved accelerated lifetime tests need to be developed. For years the industry relied on extending the qualification testing of IEC 61215 and IEC 61646 to establish confidence in its products [1]. Arbitrary extension of accelerated tests, however, may not lead to the desired result, as failure modes observed in accelerated testing should match the ones observed in the field. Therefore, detailing and quantifying failure modes and power loss from field observation is an essential part of developing lifetime service tests. The first part of this chapter is dedicated to the lifetime and gradual power decline and the second part to the various failure modes that have been documented.

3.2 Module stability and lifetime

3.2.1 Historical survey

As modules and balance-of-system components of a PV system age, they gradually lose some performance. Most commonly, PV module manufacturers guarantee against this loss by promising 80% of nameplate power after 25 years whether in a stepped or linear fashion. The decline is expressed in relative percentage, such that a module with a hypothetical absolute efficiency of 20% today would decline to 17.5% after 25 years at an annualized (relative) rate of 0.5%/year, assuming the decline is approximately linear. The decline or degradation rate has significant financial consequences, as this rate directly affects the yield and, therefore, the cash flow in future years. In financial projections, the value of a PV project after 25 years is often assumed to be negligible, yet our simple hypothetical example illustrates that such a module would present a considerable residual value after 25 years. At the time of this writing, more than 30 studies of systems older than 20 years have been reported, with some at 30 years and one even approaching 40 years. As such, there appears to be no end to the lifetime of certain PV systems. Yet other reports exist of premature decline due to manufacturing defects or poor installation quality.

For an investor, this presents a considerable quandary associated with financial risk. Therefore, it is important to understand, detail and quantify the various factors influencing the performance loss curves. Some of the authors first aggregated and analysed the public knowledge of PV lifetime measurements, which were later reanalysed [2,3]. As can be seen in the colour-coded map (A) and in the graph of studies exceeding 20 years (B) of Fig. 3.1, this topic has seen an increased interest in recent years. This chapter offers an updated analysis with a goal of providing better answers to some of the questions that remain open.

f03-01-9781782423362
Fig. 3.1 Worldwide reported degradation rates coloured by the decade of installation (A) and system reports exceeding 20 years by publication year (B).

More than 11,000 annual degradation rates have been aggregated and analysed in this section. The first observation in the histogram of Fig. 3.2A of all data (solid blue) is that the distribution is bimodal, a peculiar property that needs further discussion.

f03-02-9781782423362
Fig. 3.2 Histograms of all data, high-quality data and the median per study and system presented as the normalized frequency. The number of data points for each respective subset is given in parentheses.

As the degradation rates are impacted by a variety of factors, including (but not limited to) manufacturer, model, technology, climate, mounting, failure modes, measurement uncertainty, etc., it is important to delineate the factor causing the bimodality. Next to these hardware-related variables influencing the long-term performance assessments, statistical procedures including (but not limited to) methodology and sampling can also influence the determined rates and therefore the overall distribution. To accurately assess the status and the health of the industry in general, a representative sample is essential. As the present study is not based on a designed study but a correlational study, it is important to investigate the aggregated values in different ways to ensure the absence of unintended bias. To illustrate this, we consider this hypothetical example: 10 studies investigating the long-term performance of 10 different modules, in different climates, mounting configurations, etc., analysed together with one additional study investigating a single product, in a single location but investigating 100 modules. The aggregated distribution would be dominated by the single study but might not be representative of the product population in general. Thus, to reduce sampling bias, that is, the over- or under-representation of a group, we present the data aggregated from the literature both by including all reported measurements and by using the median of each study and system. However, even this method may not be the best methodology to find a representative sample. Consider again a hypothetical example of two studies: One study examines 10 products at high accuracy and a similar study assesses an equal number of products at very low accuracy. By using the median of each study, the sampling bias is superficially reduced but this method would equally weigh each study. However, one would probably choose to give the more accurate study more weight than the less accurate study.

Therefore, we also show a degradation rate distribution of studies that we consider to be high quality based on the following criteria: multiple measurements were taken for increased confidence; the measurement methods and calibrations were clearly described and are generally similar at each measurement point; details on the installation (disregarding proprietary considerations) are provided. Each of these analytical methodologies is imperfect, but it is interesting to note that they consistently lead to a unimodal distribution centered in the 0.5–0.8%/year range. Furthermore, the aggregation of degradation rates through survey of published information may induce sampling bias because modules or systems at high decline rates may be less likely to remain exposed in the field than those with smaller decline rates. In contrast, some studies may examine prototypes and not commercially available products without specifying. Finally, every effort was made to separate light-induced degradation (LID) from long-term performance decline. LID is a short-term performance degradation upon initial light exposure because of light-induced chemical reactions that increase the effects of defects in the semiconductor material.

Nevertheless, since we are largely limited to reporting the information provided, LID artefacts cannot be completely excluded, especially when considering data collected on thin-film technologies.

3.2.2 Nameplate rating

Many studies were found to have utilized only the nameplate rating and one measurement to determine a degradation rate slope. One-measurement studies often are a result of missing beginning-of-life measurements because they were (1) never taken or (2) because of the long life of the PV system, they can no longer be recovered. To assess the durability of the module or system, the analyst has no choice but to compare the measurement to the nameplate rating. Several implicit assumptions are made in this procedure:

1. The nameplate rating was accurate at the beginning of the PV life.

2. If an ensemble of modules is measured, the ensemble of performance starts as a point source and not as a distribution with a shape defined by a convolution of the binning practices of the manufacturer and the process control. Modules are sorted according to their measured power and allocated to specific rated power products, ‘bins’; therefore, as an example, modules from a nominal 250 W ‘bin’ may produce slightly more or less power depending on the specific manufacturer's sorting practices. Furthermore, nameplate rating constitutes a measurement by itself and several papers have discussed the uncertainties involved with the rating measurements [46]. The combination of these effects leads to the product tolerance, which has changed significantly over the past several decades.

3. Beginning-of-life LID is not considered consistently. For crystalline silicon (x-Si) this occurs on a short time scale but for some thin-film products the stabilization process may take from several months to years but is often included in the manufacturer's rating [7].

The difficulty is to take these effects into account to accurately assess PV durability, yet in many durability studies these uncertainties are often not discussed. To better gauge the uncertainty associated with one-measurement studies, Fig. 3.3 shows an extensive (yet most likely not comprehensive) graph on the history of module measurements with respect to their nameplate rating [819]. Indoor I–V measurements are indicated by red open circles and outdoor I–V data by blue open triangles. As a guide to the eye, a 10% discount line is also shown. Considerable spread exists but it appears that measurements on solar simulators tend to be closer to the nominal nameplate rating. However, the data points only go back to the mid-1990s. With the exception of a few data points in 2011, the outdoor measurements trend closer towards the nameplate rating in recent years, an effect which may be attributable to the change in manufacturers’ tolerance. At the time of the writing, most PV module manufacturers warrant tolerances of 0% to positive deviations, whereas older modules often carried equally symmetric positive/negative deviations. As discussed in the previous section, another reason for the larger outdoor deviation may be associated with the confounding effects of measurements and soiling.

f03-03-9781782423362
Fig. 3.3 Historical module I–V measurements with respect to their nameplate rating for new or almost new modules. Indoor measurements are given by the red circles and outdoor measurements are indicated by blue triangles. As a guide to the eye, the 90% value is indicated by a dashed line.

In addition LID is often contained in these data points. However, it must be emphasized that, for long-term performance, LID must also be taken into account. The 2011 study in Fig. 3.3 includes modules that had been on sun for more than 1 year; thus these modules may not truly represent beginning-of-life measurements, a fact that may explain some of the low-lying outliers. It is evident from Fig. 3.3, even if the graph is not comprehensive, that considerable deviations at the beginning of life may be expected, depending on the specific production time and tolerances. Therefore, depending on the fielded time and the specific modules, degradation calculations may be significantly impacted. These considerations should be included in an uncertainty budget. Additionally, for these reasons, one-measurement studies, which depended upon nameplate ratings, were not included in the high quality data set category in Fig. 3.2.

3.2.3 Technology trends

In this section we examine the aggregated degradation rates by technology and the evolution over the past decades. Fig. 3.4 is colour-coded by the different technologies. Multicrystalline (multi-Si) and monocrystalline (mono-Si) have been combined into one category as a significant number of studies specify x-Si but do not disclose any other details. The type of study or measurements is symbol-coded. Degradation of 1%/year and 0.5%/year are indicated by red solid and dashed lines, respectively. Each technology category is partitioned into the decade of installation wherever possible. Because of considerable data variation, boxplots representing the interquartile range and the median are overlaid over the data points.

f03-04-9781782423362
Fig. 3.4 All degradation rates (A) and the median per system and study (B) partitioned and coloured by technology and decade of installation. Studies using one measurement are indicated by an open circle and studies with two or more measurements are indicated by an open triangle. As a guide to the eye, 1%/year and 0.5%/year are indicated by solid and dashed horizontal lines, respectively.

Microcrystalline (μc-Si) is a relatively new technology; therefore, long-term performance assessments are relatively scarce yet tend to aggregate around 1%/year. In contrast, amorphous silicon (a-Si) has a history almost as long as x-Si technologies. The technology initially suffered from high degradation, and then significantly reduced degradation before an increase to above 1%/year that appears to be consistent in the past 15 years. Sufficient data are available to partition the heterojunction technology (HJT) into its own category. Similarly to μc-Si, the HIT technology is not that old; only two subcategories are available with a median degradation that appears to be consistent around 1%/year. Cadmium telluride (CdTe) shows behaviour similar to a-Si as a significant decline in degradation occurs from the first to the second decade, followed by an increase in the third decade. This increased degradation rate may be influenced by one-measurement studies and longer stabilization for some products [7]. After suffering from high initial degradation followed by the first reports of low degradation, copper indium gallium selenide (CIGS) appears to have settled for a median of around 0.5%/year.

The x-Si is the largest category and shows initial degradation of around 0.5%/year in the first decade. In subsequent decades the degradation appears to increase to slightly above 1%/year before receding to 0.5%/year in the current decade. Many of the data points in the 1990–2000 decade that showed an increase in degradation were measured using the nameplate rating. In addition, as we discussed previously, sampling bias is a serious concern. Thus, Fig. 3.4B shows the same data but uses the median of each study or system. It can be seen that most categories appear unchanged and the conclusions we stated previously are the same. Considerable change, however, can be seen in the x-Si category. The median degradation for x-Si technologies now appears to be consistently around 0.5%/year, illustrating that the higher rates reported in Fig. 3.4 for x-Si may reflect a small number of studies that reported data for thousands of modules. The mean (not shown in Fig. 3.4) for x-Si dropped from 0.99%/year for the 1980s to 0.72%/year, then was 0.78%/year and 0.77%/year for the 1990s, 2000s and 2010s, respectively.

3.2.4 Climate and mounting

Recent studies describe higher degradation rates in hotter climates compared to more moderate climates [2023], which are also reflected in the guidance provided by module manufacturers [24]. Degradation rates are influenced not only by temperature but other factors such as age, measurement uncertainties, methodologies, technology, product type and mounting. It is of considerable interest whether the aggregated data can corroborate these findings.

To avoid the confounding technology effect, Fig. 3.5 shows x-Si degradation rates (A) and the median for each study and system (B) partitioned by the number of measurements and climate zones. The climate zones are based on the Köppen–Geiger classification but aggregated into four simplified categories [25]. Statistical analysis revealed significantly higher degradation rates for studies using the nameplate rating. In addition, hotter climates show a trend towards elevated degradation rates compared to cooler climates. Mounting configuration—roof- versus rack-mounting—leading to prolonged elevated temperatures may also lead to increased degradation [23,26,27].

f03-05-9781782423362
Fig. 3.5 Aggregated degradation rates for x-Si partitioned by the number of measurements (symbol coded) and climate (colour coded). Box plots of the interquartile range with the median are overlaid on the data. All degradation rates are shown (A) and the median per study and system (B) to assess sampling bias. Studies using one measurement are indicated by an open circle and studies with two or more measurements are indicated by an open triangle. As a guide to the eye, 1%/year and 0.5%/year are indicated by solid and dashed horizontal lines, respectively.

3.2.5 Module versus system degradation

Very few studies exist where module and system performance data are investigated at the same time and thus it is not clear how module long-term performance relates to system performance. In Fig. 3.6 we partition crystalline silicon (x-Si) high-quality data into system and module data. Cumulative distribution functions for all data (solid lines) and median values per study and system (dashed lines) are shown. The number of data points is again given in brackets. The cumulative probability median is indicated by a dashed horizontal line, whereas the 0.5%/year and 1%/year degradation are indicated as vertical dashed and dash-dotted lines, respectively. At the cumulative probability median, the system curves are very similar around 0.6%/year. The degradation rates for the module data are below the systems data and show some difference depending on whether all data or median per study and system are shown, which may enable us to infer some uncertainty associated with this comparison. It follows that one would expect system degradation to be similar or slightly larger than module degradation at the median. Module and system curves start to deviate more substantially for worse-performing products. This may be consistent with a recent more detailed examination of a 20-year-old x-Si system, in which it was found that the worst-performing module was limiting the string, and, in turn, the worst-performing string was limiting the system [8]. If the modules in a given system degrade similarly, the system will degrade close to the median module; conversely, if the modules show substantial spread in their degradation behaviour, the system degradation may be significantly different from the median module behaviour.

f03-06-9781782423362
Fig. 3.6 Cumulative distribution functions for high-quality data of x-Si systems and modules. The median is indicated by a dashed horizontal line, 0.5%/year and 1%/year degradation are indicated as a dashed and dash-dotted vertical line, respectively. The number of data points for the respective subsets is given in parentheses.

3.2.6 Nonlinearities

The first challenge in long-term performance prediction is the question of linearity. Many PV technologies, especially thin-film technologies, exhibit nonlinearities at the beginning of their useful life. The initial rapid decline for amorphous silicon (a-Si) has been well documented [28]. The initial rapid decline occurs during several months before the onset of the long-term trend, as shown in Fig. 3.7. Some CdTe products show a similar behaviour [7]. The CIGS system shown in Fig. 3.7 shows a distinctly different behaviour; an initial increase during the first several months of light exposure is followed by the onset of the long-term degradation. It is evident from these examples that the initial trend is different from the long-term behaviour. In addition, if the initial phase is included in the evaluation, the long-term prediction would include a substantial, yet unintended bias.

f03-07-9781782423362
Fig. 3.7 Monthly percentage of nameplate rating not temperature corrected outdoor for thin-film technologies a-Si, CdTe and CIGS. Nonlinear tendencies can be readily observed over seasonal variations.

In addition to the beginning-of-life phase, the wear-out phase in the lifecycle of PV modules can include nonlinearities. Fig. 3.8 displays four studies examining an ensemble of modules several times during the life of a PV system, each system having been fielded for at least 20 years. The sample from the Colorado, USA study is too small to confidentially establish linearity [8]. The California, USA study shows a fairly linear behaviour for the central tendency of the modules [29]. However, the worse-performing modules display distinct signs of nonlinearity. Similar trends can be observed in the studies from Switzerland [18] and Italy [30] of modules that were exposed for 30 years.

f03-08-9781782423362
Fig. 3.8 More-than-20-year-old systems that measured an ensemble of modules multiple times during the field exposure. The modules were measured at the beginning of life, with the exception of modules for the Italian study, which were first measured after 11 years of field exposure.

3.2.7 Financial impact

On the systems level, several studies have emerged recently that partition continuous data into shorter time intervals allowing the determination of several subsequent degradation rates instead of one overall degradation rate [31,32]. This is a promising trend, as it allows delineating the properties of the degradation curve and provides more information than one overall degradation rate. Quantifying nonlinearity of degradation curves, in contrast to rates that imply linearity, can have significant impact on financial aspects of a PV project, such as the levelized cost of energy (LCOE) [33]. In summary, we may conclude that nonlinear behaviour may depend on a variety of factors, such as technology, product, climate, system load, etc., as some products will be more susceptible than others.

3.2.8 Other factors

Finally, the load under which a PV module or system is exposed may also influence the overall degradation. Crystalline Si modules were shown to have a significantly higher degradation after 20 years of field exposure when they were grid-connected through an inverter compared to open-circuit condition [34]. In contrast, for a-Si technology open-circuit conditions led to significantly higher degradation compared to under continuous load [35].

3.3 Reliability and failure modes

3.3.1 Historical survey

Service lifetime prediction, as discussed in Section 3.1, depends on developing accelerated tests that can duplicate failure modes from field observation. Information on field failure modes can be obtained empirically through site visits [36], warranty returns [37] and/or maintenance records [38]. Recently the International Energy Agency PV Power Systems Program—Task 13 published a detailed review on failures of PV modules based on literature and site visits [39]. The treatise discusses at length inspection tools, observed failures observed and new proposed test methodologies. Different failure modes are classified by safety and their different time series behaviour, eg, linear versus nonlinear. In practice it is often difficult to distinguish between different time series behaviour and estimating the impact of a failure mode for several reasons. Different products may be susceptible to different failure modes at different times in their life cycle. PV modules often display several failure modes at the same time making it difficult to isolate a certain power loss to a specific failure mode. The use environment may accelerate different failure modes differently.

Instead, a simpler strategy may be to classify failures by their modes, mechanisms, severity and detectability, as it is commonly done in other industries. Hence, failure modes are typically defined as the effect by which failures are observed [40]. In contrast, a failure mechanism is the cause that leads to failure. In Table 3.1 we summarize typically seen failures and categorize them by failure mode, failure mechanism, the severity, the onset and the detectability. Detectability frequently relies on visual inspection, yet visually observed failure modes are often subject to the inspector's knowledge and experience. A standardized protocol for visual inspection has been developed and is provided as a resource in Section 3.7 at the end of the chapter. The onset of the failure relates to different behaviour in time; it can be a gradual process or a sudden process. Yet, even this classification may be limited due to the synergistic nature of many internal and external variables. Examples of the most commonly observed failure modes and mechanisms observed in the literature, Fig. 3.9A, and visits of several hundred systems, Fig. 3.9B, will be discussed in the following section.

Table 3.1

Commonly observed PV module failure modes, mechanisms, severity, time onset and detectability

Failure modeFailure mechanismSeverityOnsetDetectability
Discolouration (EVA)Chemical alteration of cross-linkingLowGradualEvident, visual
Delamination (front side)Adhesion, loss of elastomeric propertiesLow to highGradualEvident, visual
Backsheet failureAdhesion, fractureLow to highGradual, suddenEvident, visual
Internal circuitry discolourationCorrosionLow to highGradualEvident, visual
Internal circuitry interruption (solder bonds, ribbons)Corrosion, fatigueHighSuddenVisual, requires detailed inspection
Glass breakageFractureHighSuddenEvident, visual
Cell breakageFractureLow to highSuddenEvident, visual
Hot spotsFracture, electrical mismatchLow to highGradualHidden, need IR
Burn marksFracture, electrical mismatchHighGradualEvident, visual
PIDElectrical stress, sodium migrationHighGradualHidden, need IR, EL, performance data
External circuitry disruption (j-box, bypass diode, cables)Wear, electrical stressLow to highGradualEvident, visual
Structural failuresDeformation (frame from snow load)HighSuddenEvident, visual
AR coating delaminationAdhesionLowGradualEvident, visual
SoilingWear, etching of glassLowGradualEvident, visual
LIDChemical, oxygen-boron complexLowGradualHidden, need performance data

t0010

f03-09-9781782423362
Fig. 3.9 Failure modes observed as a percentage of literature studies (A) [20] and Pareto chart of failure modes observed in a single study of multiple manufacturer (B) [41].

3.3.2 Failure modes

Discolouration

The discolouration of the ethylene vinyl acetate (EVA) encapsulant was observed in the power loss of the Carrizo Plains installation in the 1990s [42] (Fig. 3.10). Subsequently, the discolouration was detailed and appeared to have been solved [43]. However, Fig. 3.9 implies that it is still a relevant topic today, although this may be aided by the fact that discolouration is most notable by visual inspection. Discolouration typically does not cause failure, but causes decreased power output and may be a key contributor to the slow, ~ 0.5%/year degradation typically observed for PV modules. The impact is the loss of light transmission that can be readily detected in a decreased short-circuit current (Isc) [44] (Fig. 3.11).

f03-10-9781782423362
Fig. 3.10 Crystalline silicon module with significant encapsulant discolouration.
f03-11-9781782423362
Fig. 3.11 Severe delamination over an entire cell (A) [45], and large-scale delamination across multiple cells (B) [46].

Delamination

Similarly to discolouration, delamination at or near the front surface can be seen in a decreased Isc, at least in the initial stages, as an additional interface develops and loss of transmission occurs. If delamination continues and becomes more severe, moisture may ingress and lead to corrosion of the internal circuitry that can be observable in decreased fill factor (FF). Furthermore, delamination can also occur behind the x-Si cells and may manifest itself as bubbling of the backsheet or complete delamination (Fig. 3.12).

f03-12-9781782423362
Fig. 3.12 Delamination of the backsheet causing bubbles (A) and complete delamination for a frameless module (B) [26,41].

Metallization corrosion

Discolouration of the internal circuitry or corrosion is often preceded by delamination. Initially, the power loss is caused by the added interface. However, over time with increasing delamination, moisture can penetrate and lead to corrosion of the metallic circuitry. Visually corrosion is easily detectable, as shown in Fig. 3.11. Electrically, the power loss is typically associated with FF loss through increased series resistance (Fig. 3.13).

f03-13-9781782423362
Fig. 3.13 Corrosion of the metallization (the busbar and fingers) (A) [59] and interconnect (B) [60].

Potential-induced degradation

Module degradation as a function of the position in the string was first reported on high-efficiency n-type Si modules [47]. The initial confusing effect was that modules that were exposed to more positive bias displayed higher degradation. Although not fully understood at the time, the remedy was grounding and avoidance of positive bias. Only a few years later a similar effect was discovered for p-type x-Si modules placed in negative bias [48]. While more is being learned in this rapidly developing field, the cause of the effect appears to be the migration of sodium ions into the semiconductor junction [49]. This potential-induced degradation (PID) and the partial reversal of the degradation have been of significant interest in the last few years because of the rapidity and severity of the effect in some cases. Module manufacturers routinely label their products as ‘PID resistant’ or ‘PID free’, though it is not always clear what tests were performed that led to said label. Two standard test methods have been defined in IEC/TS 62804.

Cracked cells and snail trails

Another common cause of loss of Isc can be broken cells. Even though cracked cells have become more common as cells have been thinned in recent years, there are few papers written specifically about cracked cells. Publications more commonly describe ‘snail trails’ (observed when the encapsulant has a different colour near the crack than elsewhere, Fig. 3.14) and conclude that these decorate cracked cells that lead to substantial power loss [51]. However, instead of identifying the cracked cells as the problem that needs to be fixed, often the snail trails are identified as ‘the problem’. We have not seen a report that directly links the snail trails to performance loss, but note that the snail trails provide a simple way to identify cracked cells. Some PV customers wish to eliminate all cell cracks, but if a cell cracks in such a way that every fragment remains connected, there may not be power loss. In order to reduce the silicon usage, and, therefore, module cost, we need to better understand how to avoid cracks, how to test for them, and how to identify whether some cracks are acceptable [52]. Electroluminescence is a useful tool for identifying cracked cells [53] (Fig. 3.15).

f03-14-9781782423362
Fig. 3.14 Electroluminescence image of a module with many cracked cells (A) [50], and optical image of a cracked cell (B); the discolouration reveals the underlying crack.
f03-15-9781782423362
Fig. 3.15 Image showing ‘snail trails’ that decorate cracked cells [55].

Hot-spots

As individual cells are serial connected in a module substring, the current flowing through the cells is the same. If one cell is shaded, that cell is reverse-biased, dissipating power and therefore leading to heating effects. Bypass diodes are connected in parallel and in the opposite direction to the PV cells limiting the power dissipating effect (Fig. 3.16).

f03-16-9781782423362
Fig. 3.16 Hot-spots observable in infrared imaging (A, B) [61,62] and damage caused by hot-spots on front side (C) and backside of module (D) [18,41].

Partial shading is not the only situation that can lead to reverse bias of one or more cells. Mismatch between individual cells, eg, one cell producing less current than the other cells in a substring, can have a similar effect. The mismatch may increase during many years of field exposure as the individual cells degrade differently, exacerbating the effect. If the bypass diode fails to protect the substring, the heating can rapidly increase. Another ageing-related cause is the weakening connections of the internal module circuitry, such as solder bonds, through thermal cycling during field exposure. Finally, cracked cells may lead to hot-spots, as part of the cell is isolated and behaves similarly to a partially shaded cell. As the temperature difference between the hot-spot and the ambient area increases, safety becomes the primary concern. Significant damage to the front or the back of the module can occur, leading to elevated fire risk.

Glass

Glass breakage (Fig. 3.17) in thin-film modules is more common than in crystalline silicon because many thin-film products have annealed glass while most c-Si modules use tempered or heat-strengthened glass. Some arrays suffer from more glass breakage than others. Such glass breakage has been attributed to poor support structure design, inappropriate maintenance (eg, using a weed whacker to cut vegetation), poor installation and/or handling practices (eg, a small nick on the edge of a module can nucleate a crack) and finally manufacturing defects.

f03-17-9781782423362
Fig. 3.17 Cracked front cover glass for silicon (A) and thin-film (B) modules [50,54].

Frame

Bent frames are common (Fig. 3.18) in climates with snow and ice. The snow may partially melt and then refreeze along the bottom edge of the module, causing stress on the frame. Joerg Althaus has led an effort [56] to define IEC 62938 ‘Non-uniform snow load testing for PV modules’, planned for publication in 2016. This test will be valuable for those choosing modules for cold (icy) locations.

f03-18-9781782423362
Fig. 3.18 PV module frame damage because of snow load [55].

3.3.3 Standards development

Bankable PV (PV installations in which the investor has confidence in their reliable and long-term operation) benefits from international standards that address the module design, consistent manufacturing and system verification. The International PV Quality Assurance Task Force (PVQAT) was created to develop a set of universally accepted accelerated stress tests to identify adequate durability for all climates and mounting configurations [57]. PVQAT and IEC efforts to define improved PV standards are moving forward quickly to support the maturation of the industry. The challenge is to identify rapid test methods that adequately assess durability in a range of use environments while recognizing that quantitative lifetime assessment must be implemented for a specific bill of materials with a defined quality management system that specifies the variability allowed for the product implementation. PVQAT investigations will provide a basis for implementation of both climate-specific qualification tests and more quantitative service life predictions. Table 3.2 provides a summary of the module-level issues that were identified and the near- and long-term plans by PVQAT and Working Group 2 (WG2) (Modules) of IEC Technical Committee 82 (TC82) on PV.

Table 3.2

International standards efforts to address issues prioritized in Qualification Plus (Q+) [58]

PriorityCurrent statusNear-term planLong-term plan
UV durability of polymeric components (encapsulants, backsheets, connectors and junction boxes) and as a requirement for insulation materials (see Q + component tests 1–4)Recent and planned changes will add more UV and other testing of polymeric materials(1) Submit 62788-7-2;
(2) Submit amendment to IEC 61730 referencing IEC 62788-7-2 for materials weathering conditions;
(3) Submit additional IEC 62788 parts; publish in 2016
Define climate-specific UV exposures
Bypass diode and j-box thermal test (see Q + component test 5)Multiple documents are in progress to address thermal endurance, cycling, and runawayFive actions are described in text reference [55]Define climate-specific diode tests
Electrical failures within the module (see Q + module test 1)Ribbon interconnects can be tested with cyclic mechanical loading, but thermal cycling (TC) is still neededDefine climate-specific thermal-cycle tests and implementation within QMS
Power loss from cracked cells (see Q + module test 2)IEC 62782 TS: Cyclic (dynamic) mechanical load testing and IEC 62759-1 PV modules— Transportation testing were publishedSubmit amendment to IEC 61215 to add 1000 cycles from IEC 62782 before TC and humidity freezeDefine climate-specific tests for cracked cells, if needed
Susceptibility to PID (see Q + module test 3)IEC/TS 62804-1 Test methods for detection of PID—Part 1 Crystalline Silicon has been publishedDefine standard labels for PID susceptibilityUnderstand relationship between degradation in the test and in the field; Identify quick tests for screening cells and encapsulant materials
Susceptibility to hot spot degradation (see Q + module test 4)Revised hot-spot test is included in 2016 version of IEC 61215Define improved hot-spot test for thin-film modules
Improve confidence in Quality Management System (QMS) (see Q + description of sampling and QMS)IEC/TS 62941 Guideline for increased confidence in PV module design qualification and type approval was publishedImplement through IECREFacilitate adoption and assess value of extending to quantitative assessment
Structural failure from snow and ice (based on experience in Europe and New England)IEC 62938 Non-uniform snow load testing for PV modules Committee Draft has been reviewedComplete IEC 62938; publish in 2017Encourage use of the snow load test to differentiate modules
Faster degradation in hot climatesDelamination, encapsulant discolouration, and thermal fatigue are documented to increase with high temperatureDefine difference in use environmentUse IEC 62892 to implement comprehensive tests at higher temperatures
Assessment of system functionalityDrafts completed for both energy and capacity testsPublication of IEC 61724-2 and IEC 61724-3 in 2016Implement as part of IECRE

t0015

3.4 Conclusion and future trends

Today, the PV industry is fairly knowledgeable about the most common causes of failures in the field, but a complete understanding of the acceptable process and design windows is challenging, especially as companies may make changes to the bill of materials frequently. PV reliability standards are maturing by:

 addressing failure mechanisms that have recently become problematic;

 differentiating the durability of PV module designs as a function of the use environment and

 improving manufacturing consistency toward a consistent implementation of product design, moving toward inclusion of service life prediction in the most mature quality management systems.

Today's most critical issues will be addressed by standards completed in 2016. Quantitative service life predictions will become practical when the industry is mature enough for product designs to stabilize enough to allow time for testing and model development and validation.

3.5 Resources

3.5.1 Visual inspection sheet

u03-01p1-9781782423362u03-01p2-9781782423362u03-01p3-9781782423362u03-01p4-9781782423362u03-01p5-9781782423362u03-01p6-9781782423362

References

[1] International Electrotechnical Commission. Crystalline Silicon Terrestrial Photovoltaic (PV) Modules—Design Qualification and Type Approval. first ed. Geneva: IEC Central Office; 2016 IEC 61215.

[2] Jordan D.C., Kurtz S.R. Photovoltaic degradation rates—an analytical review. Prog. Photovolt. Res. Appl. 2013;21(1):12–29. doi:10.1002/pip.1182.

[3] J. Newmiller, G.L. DNV, White paper on photovoltaic module degradation, RANA-WP-03-A, 2015.

[4] Ransome S. Modelling inaccuracies of PV energy yield simulations. In: Proceedings of the 33rd PV Specialists Conference, San Diego, CA, USA; 2008:1–5. doi:10.1109/PVSC.2008.4922864.

[5] Dirnberger D., Müller B., Reise C. PV module energy rating: opportunities and limitations. Prog. Photovolt. Res. Appl. 2015;1:doi:10.1002/pip.2618.

[6] Dirnberger D. Uncertainty in PV module measurement—part II: verification of rated power and stability problems. J. Photovolt. 2014;4(3):991–1007.

[7] Ngan L., Strevel N., Passow K., Panchula A.F., Jordan D. Performance characterization of cadmium telluride modules validated by utility-scale and test systems. In: Proceedings of the 40th IEEE Photovoltaic Specialists Conference, Denver, CO, USA; 2014:1957–1962.

[8] Jordan D.C., Sekulic B., Marion B., Kurtz S.R. Performance and aging of a 20-year-old silicon PV system. J. Photovolt. 2015;5(3):744–751. doi:10.1109/JPHOTOV.2015.2396360.

[9] TamizhMani M.G., Kuitche J., Mikonowicz A. Nameplate, datasheet and sampling requirements for photovoltaic modules. In: Solar ABC Stakeholder Workshop, Las Vegas, 2014; 2014.

[10] Detrick A., Kimber A., Mitchell L. Performance evaluation standards for photovoltaic modules and systems, performance evaluation standards for photovoltaic modules & systems. In: Proceedings of the 31th IEEE Photovoltaic Specialists Conference, Lake Buena Vista, FL; 2005:1581–1586.

[11] García M., Marroyo L., Lorenzo E., Marcos J., Pérez M. STC power for 15 MW of PV comparing nameplate, initial power and power after 4 years. Prog. Photovolt. Res. Appl. 2012;22(2):242–247.

[12] Leloux J., Narvarte L., Trebosc D. Performance analysis of 10,000 residential PV systems in France and Belgium. In: Proceedings of the 26th European Photovoltaic Solar Energy Conference, Hamburg, Germany; 2011:3939–3946.

[13] Oozeki T., Yamada T., Otani K., Takashima T., Kato K. Performance trends in grid-connected photovoltaic systems for public and industrial use in Japan. Prog. Photovolt. Res. Appl. 2010;18(8):596–602. doi:10.1002/pip.992.

[14] Reis A.M., Coleman N.T., Marshall M.W., Lehman P.A., Chamberlin C.E. Comparison of PV module performance before and after 11-years of field exposure. In: Proceedings of the 29th PV Specialists Conference, New Orleans, LA, USA; 2002:1432–1435. doi:10.1109/PVSC.2002.1190878.

[15] Coiante O., Castello S. PV module performance after ten years of continuous outdoor operations. In: ISES 1491—Solar World Congress; 1991.

[16] Granata J.E., Boyson W.E., Kratochvil J.A., Quintana M.A. Long-term performance and reliability assessment of 8 PV arrays at Sandia National Laboratories. In: Proceedings of the 34th PV Specialists Conference, Philadelphia, PA, USA; 2009:1486–1491. doi:10.1109/PVSC.2009.5411336.

[17] Sanchez-Friera P., Piliougine M., Pelaez J., Carretero J., Sidrach de Cardona M. Analysis of degradation mechanisms of crystalline silicon PV modules after 12 years of operation in Southern Europe. Prog. Photovoltaics: Res. Appl. 2011;19(6):658–666. doi:10.1002/pip.1083.

[18] Friesen T., Chianese D., Realini A., Friesen G., Burà E., Virtuani A., Strepparava D., Meoli R. TISO 10 kW: 30 Years Experience with a PV Plant. In: Proceedings of the 27th European Photovoltaic Solar Energy Conference, Frankfurt, Germany; 2012:3125–3131.

[19] Hamzavy B.T., Bradley A.Z. Safety and performance analysis of a commercial photovoltaic installation. Proc. SPIE. 2013;8825:doi:10.1117/12.2023406.

[20] Jordan D.C., Wohlgemuth J.H., Kurtz S.R. Technology and climate trends in PV module degradation. In: Proceedings of the 27th European Photovoltaic Solar Energy Conference, Frankfurt, Germany; 2012:3118–3124.

[21] Singh J., Belmont J., TamizhMani G. Degradation analysis of 1900 PV modules in a hot-dry climate: results after 12 to 18 years of field exposure. In: Proceedings of the 39th IEEE Photovoltaic Specialists Conference, Tampa, FL, USA; 2013:3270–3275.

[22] Dubey R., Chattopadhyay S., Kuthanazhi V., John J.J., Solanki C.S., Kottantharayil A., Arora B.M., Narasimhan K.L., Vasi J.M., Kumar A., Sastry O.S. Performance degradation in field-aged crystalline silicon PV modules in different indian climatic conditions. In: Proceedings of the 40th IEEE Photovoltaic Specialists Conference, Denver, CO, USA; 2014:3182–3187.

[23] Ying Ye J., Reindl T., Aberle A.G., Walsh T.M. Performance Degradation of Various PV Module Technologies in Tropical Singapore. J. Photovolt. 2014;4(5):1288–1294.

[24] Strevel N., Trippel L., Gloeckler M. Performance characterization and Superior energy yield of first solar PV power plants in high-temperature conditions. In: Photovoltaics International; 2012.

[25] Desert: BWh, BWk; Hot & Humid: Af, Am, As, Aw, Cwa, Cfa; Moderate: Cfb, Csa, Csb, BSh, BSk; Snow: Dfa, Dfb, Dfc, ET.

[26] Yedidi K., Tatapudi S., Mallineni J., Knisely B., Kutiche J., TamizhMani G. Failure and degradation modes and rates of PV modules in a hot-dry climate: results after 16 years of field exposure. In: Proceedings of the 40th IEEE Photovoltaic Specialists Conference, Denver, CO, USA; 2014:3245–3247.

[27] Mikofski M.A., Kavulak D.F.J., Okawa D., Shen Y.-C., Terao A., Anderson M., Caldwell S., Kim D., Boitnott N., Castro J., Smith L.A.L., Lacerda R., Benjamin D., Hasselbrink Jr. E.F. PVlife: an integrated model for predicting PV performance degradation over 25 + years. In: Proceedings of the 38th IEEE Photovoltaic Specialists Conference, Austin, TX USA; 2012:1744–1749.

[28] Staebler D.L., Wronski C.R. Optically induced conductivity changes in discharge-produced hydrogenated amorphous silicon. J. Appl. Phys. 1980;51(6):3262–3268.

[29] Chamberlin C.E., Rocheleau M.A., Marshall M.W., Reis A.M., Coleman N.T., Lehman P.A. Comparison of PV module performance before and after 11 and 20 years of field exposure. In: Proceedings of the 37th IEEE Photovoltaic Specialists Conference, Seattle, WA USA; 2011:101–105.

[30] L. Abenante, F. De Lia, S. Castello, Long-term performance degradation of c-Si photovoltaic modules and Strings, in: Proceedings of the 25th European photovoltaic solar energy conference, Valencia, Spain, 4023–4026

[31] Kyprianou A., Phinikarides A., Makrides G., Georghiou G.E. Robust principal component analysis for computing the degradation rates of different photovoltaic systems. In: 2014:2939–2942. Proceedings of the 29th European Photovoltaic Solar Energy Conference, Amsterdam, Netherlands.

[32] Belluardo G., Ingenhoven P., Sparber W., Wagner J., Weihs P., Moser D. Novel method for the improvement in the evaluation of outdoor performance loss rate in different PV technologies and comparison with two other methods. Sol. Energ. 2015;117:139–152.

[33] D.C. Jordan, S.R. Kurtz , K.T. VanSant , J. Newmiller, Compendium of photovoltaic degradation rates, Progress in PV, 2016, http://dx.doi.org/10.1002/pip.2744/

[34] Skoczek A., Sample T., Dunlop E.D. The results of performance measurements of field-aged crystalline silicon photovoltaic modules. Prog. Photovolt. Res. Appl. 2009;17:227–240. doi:10.1002/pip.874.

[35] Fanni L., Pola I., Burà E., Friesen T., Chianese D. Investigation of annealing and degradation effects on a-Si PV modules in real operating conditions. In: Proceedings of 24th European Photovoltaic Solar Energy Conference, Hamburg, Germany; 2009:3596–3599. doi:10.4229/24thEUPVSEC2009-4AV.3.92.

[36] DeGraaff D., Lacerda R., Campeau Z., Xie Z. How do qualified modules fail—what is the root cause? In: International PV Module Quality Assurance Forum, San Francisco, CA; 2011.

[37] Wohlgemuth J.H., Cunningham D.W., Nguyen A.M., Miller J. Long term reliability of PV modules. In: 20th European PV Solar Energy Conference, Barcelona, Spain; 2005:1942–1948.

[38] Golnas A. PV system reliability: an operator's perspective. IEEE J. Photovolt. 2013;3(1):416–421. doi:10.1109/JPHOTOV.2012.2215015.

[39] M. Köntges, S. Kurtz, C. Packard, U. Jahn, K.A. Berger, K. Kato, T. Friesen, H. Liu, M. Van Iseghem, Performance and reliability of photovoltaic systems, subtask 3.2: review of failures of photovoltaic modules, International Energy Agency, Report IEA-PVPS T13-01:2014, 2014.

[40] SAE Standard SAE J1739, 2002.

[41] DeGraaff D., Lacerda R., Campeau Z. Degradation mechanisms in Si module technologies observed in the field; their analysis and statistics. In: PV Module Reliability Workshop, Golden, CO, USA; 2011.

[42] Rosenthal A.L., Lane G.C. Field test results for the 6 MW Carrizo solar PV power plant. Solar Cells. 1991;30(1–4):563–571. doi:10.1016/0379-6787(91)90088-7.

[43] W.W. Holey, S.C. Agro, Advanced EVA-based encapsulants, NREL/SR-520-25296, 1998.

[44] Smith R.M., Jordan D.C., Kurtz S.R., Outdoor P.V. Module degradation of current–voltage parameters. In: World Renewable Energy Forum, Denver, CO, USA; 2012.

[45] Skoczek A., Sample T., Dunlop E.D., Ossenbrink H.A. Electrical performance results from long-term outdoor weathered modules. In: 22nd European Photovoltaic Solar Energy Conference, Milan, Italy; 2007:2458–2466.

[46] Alonso-Abella M., Chenlo F., Alonso A., González D. Toledo PV plant 1 MWp—20 years of operation. In: Proceedings of the 29th European Photovoltaic Solar Energy Conference, Amsterdam, Netherlands; 2014:2728–2733.

[47] Swanson R., Cudzinovic M., DeCeuster D., Desai V., Jürgens J., Kaminar N., Mulligan W., Rodrigues-Barbarosa L., Rose D., Smith D., Terao A., Wilson K. The surface polarization effect in high efficiency silicon solar cells. In: Proceedings of the 15th International Photovoltaic Science and Engineering Conference, Shanghai, China; 2005.

[48] Pingel S., Frank O., Winkler M., Daryan S., Geipel T., Hoehne H., Berghold J. Potentially induced degradation of solar cells and panels. In: Proceedings of the 35th IEEE PV Specialists Conference, Honolulu, HI, USA; 2010:2817–2822.

[49] Hacke P., Terwilliger K., Smith R., Glick S., Pankow J., Kempe M., Bennett S.K.I., Kloos M. System voltage potential-induced degradation mechanisms in PV modules and methods for test. In: Proceedings of the 37th IEEE PV Specialists Conference, Seattle, WA, USA; 2011:814–820.

[50] Wohlgemuth J.H., Kurtz S.R. PV module reliability: how can we improve it? In: 30th European Photovoltaic Solar Energy Conference, Hamburg, Germany; 2015.

[51] Dolara A., Leva S., Manzolini G., Ogliari E. Investigation on performance decay on photovoltaic modules: snail trails and cell microcracks. IEEE Journal of Phtovoltaics. 2014;4(5):1204–1211.

[52] Köntges M., Kajari-Schroder S., Kunze I. Crack statistic for wafer-based silicon solar cell modules in the field measured by UV fluorescence. IEEE J. Photovolt. 2013;3:95–101.

[53] Mathiak G., Beniaminova J., Pohl L., Sommer J., Herrmann W., Reil F. Quantitative electroluminescence imaging applied to hail impact damages. In: 29th European Photovoltaic Solar Energy Conference, Amsterdam, Netherlands; 2014:3285–3287.

[54] Quintana M.A., King D.L., McMahon T.J., Osterwald C.R. Commonly observed degradation in field-aged photovoltaic modules. In: Proceedings of the 22nd IEEE Photovoltaic Specialists Conference, Las Vegas, NV, USA; 2002:1436–1439.

[55] Kurtz S., Sample T., Wohlgemuth J., Zhou W., Bosco N., Althaus J., Phillips N., Deceglie M., Flueckiger C., Hacke P., Miller D., Kempe M., Yamamichi M., Kondo M. In: in: Proceedings of the 42nd IEEE Photovoltaic Specialists Conference, New Orleans, LA, USA; 2015.

[56] Reil F., Mathiak G., Strohkendl K., Raubach S., Schloth C., Wangenheim B.V. Experimental testing of PV modules under inhomogeneous snow loads. In: 27th European Photovoltaic Solar Energy Conference, Frankfurt, Germany; 2012:3414–3417.

[57] http://www.pvqat.org/.

[58] S. Kurtz, J. Wohlgemuth, M. Kempe, N. Bosco, P. Hacke, D. Jordan, D.C. Miller, T.J. Silverman, N. Phillips, T. Earnest, R. Romero, Photovoltaic Module Qualification Plus Testing, NREL/TP-5200-60950, 2013.

[59] Wohlgemuth J., Silverman T., Miller D.C., McNutt P., Kempe M., Deceglie M. Evaluation of PV module field performance. In: Proceedings of the 42nd IEEE Photovoltaic Specialists Conference, New Orleans, LA, USA; 2015.

[60] Packard C.E., Wohlgemuth J.H., Kurtz S.R., Jahn U., Berger K., Friesen T., Koentges M. Fielded PV module condition. In: 27th European Photovoltaic Solar Energy Conference, Frankfurt, Germany; 2012.

[61] García M., Marroyo L., Lorenzo E., Marcos J., Pérez M. Observed degradation in photovoltaic plants affected by hot-spots. Prog. Photovolt. Res. Appl. 2013;22:1292–1301. doi:10.1002/pip.2393.

[62] Kato K. PVRessQ!. In: PV Module Failures Observed in the Field, PV Module Reliability Workshop, Golden, CO; 2012.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.222.118.159