Sunteți pe pagina 1din 75

A heating and cooling load benchmark

model for a Dutch office building using an


ANN approach

Master thesis Msc Business Information Management

Rotterdam School of Management, Erasmus University

Author: Damiën Horsten

Student number: 328815

Thesis coach: Dr Tobias Brandt

Co-reader: Professor Wolfgang Ketter

Submission date: 15-08-2017

1
Preface
The copyright of the master thesis rests with the author. The author is responsible for its

contents. RSM is only responsible for the educational coaching and cannot be held liable for

the content.

2
Abstract
In this thesis, we develop a data-driven model to benchmark the heating and cooling load for

a Dutch office building. This poses an alternative to the default option of building a physical

model which is both time-consuming and expensive. For model training and testing, we use

one years’ worth of data that was recorded in one-hour intervals over 2016. A significant

amount of missing data complicates more traditional time series analyses.

The object of research is located in the Netherlands, making it an interesting case study

because of the country’s mixed climate: most of the existing literature concerns prediction

models for either heating or (most often) cooling. The building is equipped with several

heating and cooling subsystems as well as some energy saving features. Apart from a large

amount of heating/cooling related parameters that were monitored by the building

management system, we also have access to some figures regarding electricity consumption.

We use this to create an approximation of occupancy which, as literature describes, is one of

the main drivers of building heating/cooling load. For the weather data we use the 2016 (and

2015) datasets that are made available by the KNMI. We use both the hourly and the daily

datasets for the corresponding models. Partly based on the literature review we select a

scarce number of explaining variables. Specifically, we achieved the best results including as

weather variables mean and maximum temperature, global solar radiation, wind speed and

relative humidity.

We use an Artificial Neural Network (ANN) to construct the model and compare its

performance to a simpler linear regression model. Also, separate models are built for heating

and cooling and for hourly and daily predictions, so that we end up with 8 models. Features

are selected based on physical principles. Lagged variables are used to account for thermal

inertia of the building. Apart from tuning the ANN, some techniques are introduced to

improve the results. We apply cross-validation to prevent overfitting to a specific training set

and for the daily heating model we apply zero-inflation post-processing.

The final results for the daily neural network models are promising. The results for the daily

heating and cooling models are comfortably within the bounds suggested by ASHRAE

guidelines. Also, both models outperform the existing physical model. The hourly models

are unable to achieve a level of accuracy that can be accepted under ASHRAE standards. An

3
attempt to validate the results further is made by using data on the last 45 days of 2015 as a

validation set. Because this period comprises entirely of a winter period the heating model

performs very well but the cooling model struggles.

The neural network models without exception significantly outperform the linear regression

models, leading us to conclude that the underlying physical principles are too complicated to

be (efficiently) modelled using linear regression. Also, the addition of the occupancy-related

parameters to the model without exception results in a significant improvement of prediction

accuracy. We conclude with some directions for further research that we feel might improve

results, but that we were unable to perform ourselves due to time limitations.

4
Table of Contents
Preface .................................................................................................................................................... 2

Abstract .................................................................................................................................................. 3

1. Introduction................................................................................................................................... 7

2. Literature review I: types of models, feature selection, load prediction and relevance ... 10

2.1 Physical vs. data-driven models ....................................................................................... 10

2.2 Feature selection ................................................................................................................. 13

2.3 Load prediction research ................................................................................................... 14

2.4 Relevance ............................................................................................................................. 20

3. Literature review II: ANN’s & ANN parameters .................................................................. 22

3.1 ANN’s .................................................................................................................................. 22

3.2 ANN parameters ................................................................................................................ 24

4. Research objective and conceptual model .............................................................................. 26

4.1 Research objective .............................................................................................................. 26

4.2 Research questions ............................................................................................................. 26

4.3. Conceptual model .............................................................................................................. 27

5. Case study ................................................................................................................................... 28

5.1 Building and subsystems................................................................................................... 28

5.2 Air handling units: heat recovery system ....................................................................... 29

5.3 Floor heating/cooling ......................................................................................................... 30

5.4 Residual electricity consumption ..................................................................................... 30

6. Data description.......................................................................................................................... 32

6.1 Missing values..................................................................................................................... 42

7. Methodology ............................................................................................................................... 45

7.1 Data preparation ................................................................................................................. 45

7.2 Seasonality ........................................................................................................................... 46

5
7.3 Feature selection ................................................................................................................. 46

7.4 Main models ........................................................................................................................ 49

7.4.1 Linear regression model ............................................................................................ 49

7.4.2 Neural Network model.............................................................................................. 50

7.5 Post-processing: ‘Zero-inflation’ ...................................................................................... 50

7.6 Performance measures ....................................................................................................... 51

8. Results .......................................................................................................................................... 52

8.1 Zero-inflation ...................................................................................................................... 57

8.2 Additional validation: 2015 data ...................................................................................... 58

8.3 Computational load ........................................................................................................... 59

9. Discussion .................................................................................................................................... 61

10. Theoretical and managerial contribution ............................................................................ 63

11. Limitations and suggestions for further research .............................................................. 64

11.1 Suggestions for further research....................................................................................... 64

Literature list ....................................................................................................................................... 67

Appendix A: AHU, Toutdoor, Toutdoor setpoint data descriptives .......................................... 71

Appendix B: Neural network visualisation and description for daily cooling model ............. 73

6
1. Introduction
For a variety of reasons, modelling the energy performance of buildings – whether it be for

instance cooling/heating load, electricity consumption or heat loss coefficient – is potentially

valuable to companies. For this and other reasons, it has become a popular research subject

over the past decades, and this interest appears to be growing still.

The most obvious business case one can think of would likely be reducing building energy

consumption by using such models. This in itself is a very relevant theme: according to

Pérez-Lombard et al. (2007), the contribution of buildings to total final energy consumption

in developed countries is estimated to lie between 20% and 40%. In Europe, 40% of total

energy use can be accounted to buildings (Zhao & Magoulès, 2012). By 2015, however, Ürge-

Vorsatz et al. (2015) estimated the building sector to be the world’s largest consumer of

energy, accounting for 32% of final energy consumption. Consequently, the sector is

responsible for roughly 33% of all greenhouse gas emissions. According to Fan et al. (2017),

building energy consumption is for the largest part – ca. 50% – attributable to heating,

ventilation and air conditioning (HVAC) systems. This in turn leads to these types of services

offering the greatest saving potential. A decade earlier, Pérez-Lombard et al. (2007) estimated

that HVAC systems combined accounted for 50% of building energy consumption and for

20% of total consumption in the USA. Although these studies are not 1-on-1 comparable,

they indicate the proportion of HVAC systems to total final energy consumption may be

rising. Also, these figures give a sense of just how much energy buildings, and HVAC

systems in particular, consume.

For this reason, and as mentioned earlier, a great amount of research has been conducted on

the topic of building energy consumption. Most of it is predictive in nature (e.g. Datta et al.

(2000), Ben-Nakhi & Mahmoud (2004), Beto & Fiorelli (2008), Taieb & Hyndman (2013), Fan

et al. (2014) and Fan et al. (2017)). Using models to forecast, for instance, the electricity

consumption of a building for the next 24 hours has benefits, the most notable of which, at

least for the business side is the ability to purchase the electricity at a better rate when the

estimated amount required becomes known earlier.

Prediction of heating or cooling load – defined by Merriam-Webster as ‘the quantity of

heat/cold per unit time that must be supplied to maintain the temperature in a building or

7
portion of a building at a given level’ – is equally common. Fan et al. (2017) state that short

term prediction of building cooling load is useful for among others fault detection and

diagnosis, demand-side management and control-optimization. The same holds for heating

load, but this specific research was conducted in a relatively hot part of the world

(Hongkong). Regarding control strategy optimization, these authors, referring to Ben-Nakhi

& Mahmoud (2004) indicate that operating costs can be reduced whist operating flexibility

can be increased. Regarding demand-side management, a lot of research has been performed

on the topic of load shifting to decrease costs, especially in the intersection of buildings and

smart grids (e.g. Xue et al., 2014).

In this thesis, we will be developing a heating/cooling load benchmark model for a medium-

size office building in the Netherlands. Two factors differentiate the present study from

existing research: first, most available studies were conducted in Asia, which means that the

usually focus on cooling load. There are also studies on heating load present, but

surprisingly not much research appears to have been conducted in mixed-climate

environments like the Netherlands. Second, contrary to most other research, we will not be

focusing on forecasting ability. Rather, we will attempt to create a model that can function as
a benchmark. The purposes of this are numerous, as shall be illustrated.

This thesis has been written in collaboration with DWA. This Dutch HVAC consultancy firm

was founded in 1986 and consists of a staff of 110 professionals from a variety of disciplines.

The company focuses on sustainability matters and works for – among others –

governments, commercial companies, education and healthcare institutions and housing

corporations, especially where the environment, climate and energy are concerned. DWA

operates from four locations in the Netherlands, with their headquarters situated in

Bodegraven.

The remainder of this thesis is organised as follows: Chapters 2 and 3 give an overview of the

existing research on the topics of modelling energy consumption and ANN’s. Here, the

business benefits of the benchmark model that is developed will also be clarified. Chapter 4

contains the research objective and the conceptual model. In Chapter 5, the case study of the

Dutch office building is presented. In Chapter 6, the dataset that was made available to us by

DWA is discussed. In Chapter 7, a description of the methodology is given. In Chapter 8, we

8
present the results, which are thereafter discussed in Chapter 9. We conclude with

explaining the theoretical and managerial contribution of this study in Chapter 10 and with

some final remarks, limitations and recommendations for future research in Chapter 11.

9
2. Literature review I: types of models, feature selection, load
prediction and relevance
A large amount of research has been performed on the topic of building heating and cooling

load prediction. Common topics are finding the most important drivers thereof as well as the

optimal model type and design. In this chapter we will discuss research on the theme of

predicting energy load for buildings. Also, we will dedicate a section to discuss the relevance

of this research.

Given the ambiguity1 that exists regarding the meaning of the word ‘prediction’, we want to
give a clarification of the meaning we intend (and indeed, this field of research appears to
intend) with it. Strictly speaking, one can argue that predicting something by definition
refers to determining something of which the value is not yet known. In temporal prediction
(for time series data, like in the present study) this would mean that in order to predict
something, an unknown value in the future needs to be calculated. This is not, however,
what the majority of the studies we encountered describes with ‘prediction’. Although some
studies try to predict into the future (estimate the value of Y on t+1 with the value of X on
time t), mostly, the term prediction is being used to refer to a model that, using a training
and testing approach, tries to examine the value of Y on time t, using factors on time t.
Therefore, we will conform to this meaning and strictly distinguish the
prediction/benchmark model we are building from forecasting models which,
unquestionably, refer to estimating future values.

2.1 Physical vs. data-driven models


Generally speaking, there are two types of models that exist for the explanation and

prediction of building energy consumption (Fan et al., 2017). The first type, the so-called

white-box models are physical models. Although accurate results can be obtained using such

models, their main drawback is the complexity of generating them. Information on building

charateristics - such as materials and insulation used but also for instance air penetration – to

the finest level are required for these types of models to perform well (Wisse, 2017).

1E.g. https://stats.stackexchange.com/questions/65287/difference-between-forecast-and-prediction;
http://www.analyticbridge.datasciencecentral.com/forum/topics/difference-between-prediction;
https://www.linkedin.com/pulse/difference-between-forecast-prediction-lakshminarayanan-t-p.

10
Lai et al. (2008) describe how physical models are time-consuming to build, as ‘each

subsystem, each physical parameter’ needs to be carefully defined.

On the other hand, there are the data-driven models. These are regularly referred to as grey

or black-box models. These use data that was obtained during operation to model the

relationships between all sorts of operational variables and the heating/cooling/energy load

that needs to be explained or forecast. Their main advantage is the efficiency with which they

can be built, compared to physical models. Using advanced techniques such as ANN’s or

SVM’s, a high level of accuracy can be obtained using these models also.

Zhao & Magoulès (2012) state that in order to come up with a data-driven model, there are

three suitable methods for performing the analyses and the predictions:

A. Statistical methods. This is the classical way of developing a regression model that

links the independent variables to the dependent variable (energy consumption);

B. Support Vector Machines (SVM’s). Also perform well when dealing with non-linear

problems. A drawback to using SVM’s is that the training times can increase very

quickly once the datasets become larger (Meyer, 2017);

C. Artificial Neural Networks (ANN’s). Popular in this particular field of research


because of their good performance at solving non-linear problems.

SVM’s are binary classifiers that try to find the optimal hyperplane separating the two

classes, thereby maximizing the space between the closest points. The name is derived from

the fact that the points lying on the boundaries are called support vectors. (Meyer, 2017)

ANN’s are, as the name suggest, loosely based on the way our brains work, in terms of their

structure. They are organized in layers (input layer, hidden layers and output layer) and the

nodes within layers are interconnected. There are many indications for non-linear techniques

being able to outperform linear ones such as multiple linear regression and ARMA/ARIMA

(Fan et al., 2017). While it is possible to include non-linear relationships in regression models,

this needs to be done manually for such models. SVM’s or ANN’s, on the other hand, can by

themselves recognize and include non-linear relationships during model training.

11
It is important to note that, although these are the most used methods, many others exist.

Taieb & Hyndman (2013) for instance very successfully applied gradient boosting machines

and extreme gradient boosting to a load forecasting problem.

As Neto & Fiorelli (2008) describe in their paper, ANN’s are, contrary to more traditional

methods, capable of ‘learning the rule that controls a physical phenomenon under

consideration from previously known situations’. This is precisely why they are so popular
in this field of research.

Neto & Fiorelli (2008) made a comparison between a physical principles (EnergyPlus) model

and a data driven model. Their object of research was the administration building of the

University of São Paulo, covering about 3.000 m2 and with a building population of ca. 1.000

employees. Lowering of operational costs can be realised, according to these authors, in two

ways: by evaluating the energy end-users and implementing actions to reduce the

consumption, or by lowering the penalties that are often imposed by electricity operators

when peak loads exceed a certain, often contractually determined threshold. EnergyPlus is a

building simulation software that allows for implementation of geometry, materials, internal

loads, HVAC characteristics and so on. It supports daily and annual simulations. The

parameters of the simulation are determined by the user through weather parameter

profiles. Both the physical model as well as the basic ANN model were able to achieve a

good fit in 80% of the cases. The authors showed that dry-bulb temperature alone is a very

good predictor; adding other parameters (i.e. humidity and solar radiation) improved the

models, but not drastically so. The authors further show that creating separate models for

working days and non-working days can significantly improve results. An important lesson

from this paper is just how much uncertainty is being introduced by occupants’ behaviour.

They refer with this mainly to responses of occupants to a decreased thermal comfort, for

instance by opening the windows. Their most optimized neural network was able to achieve

an accuracy of about 90%.

12
2.2 Feature selection
Feature selection refers to the process of selecting relevant predictors for a model. Accuracy

and computation related benefits aside, one important reason for using feature extraction is

that overfitting risks are decreased significantly, compared to just using all available inputs

and lags (Fan et al., 2017). As Fay et al. (2003) discuss, feature selection is one of the most

important steps in building prediction models due to the poor generalisation that results

from selecting the wrong features. These authors also note that dimensionality reduction can

benefit neural network performance.

In the case of ANN’s, feature selection might be less of a necessity than for other types of

models. To quote Kalogirou et al. (1997): ‘[ANN’s] seem to simply ignore excess data that are

of minimal significance and concentrate instead on the more important inputs’. This has been

confirmed by other authors also. Nevertheless, feature selection is almost without exception

applied to ANN models also and usually does improve performance. Also, it has other
benefits such as a decrease in computational load.

Regarding their feature selection process, Fan et al. (2014) describe two main methods,

namely feature reconstruction methods such as principal component analysis and embedded

methods, such as RFE. The authors utilised a sophisticated ‘data driven input selection

process’, based on RFE to select relevant features. They further applied a clustering analysis

to detect underlying patterns in the data.

The most popular methods of relevant feature selection are engineering knowledge (e.g.

outside temperature, occupancy) and simple statistical methods. However, more advanced

‘feature extraction’ methods have also been applied. Fan et al. (2017) used and compared

four types of feature extraction, namely engineering, statistical, structural and deep learning.

Structural in this context refers to the structural relationships that exist within the data over a

certain time period. The unsupervised deep learning approach applied by these authors in

this respect stood out in both performance as well as simplicity. Very little domain

knowledge was required for this method to extract relevant features.

13
2.3 Load prediction research
Yang & Becerik-Gerber (2017) estimate that 40 percent of energy consumption in the USA is

attributable to buildings, almost half of that to commercial buildings and 40 percent of that to

HVAC systems in commercial buildings. Improving this figure, the authors explain, can be

realised by improving the thermal properties of buildings (e.g. insulation) or by better

controlling the HVAC: oftentimes, these systems are not ran in an energy-efficient manner.

An extreme example of this would be the heating and cooling systems continuously battling

each other to achieve a pleasant environment. A more real world example is given by Bauer

& Scartezzini (1998): some control systems heat a building up in the morning because of the

low outdoor temperature, only to cool it down a gain in the afternoon due to a high outdoor

temperature. This is called the pendulum effect and, although sometimes required for

thermal comfort, this can partly be handled more efficiently. Such inefficiencies can be

realised without losing thermal comfort and far more cost-efficient than by the first, physical

solution of improving the thermal properties of the building. The central theme in the Yang

& Becerik-Gerber (2017) paper are state transitions, referring to the switches such as arrival,

intermittent absence and departure. The authors measured such transitions through setpoint

analysis. Using an Enhanced Variable Neighbourhood Algorithm based on these transitions,

they were able to optimize (simulated) energy efficiency by between 10 and 28 percent.

Zhao & Magoulès (2012) state that the most suitable and popular methods to perform energy

load analyses are Support Vector Machines (SVM’s) and Artificial Neural Networks

(ANN’s). However, more conservative statistical (regression) methods have been used to

predict energy consumption with relatively simple input variables. This has been done by,

among others, Bauer & Scartezzini (1998) and Cho et al. (2004). Bauer & Scartezzini (1998)

indicate that – of course depending on properties of the building envelope (glazing

especially) and the climate, during the summer, solar irradiation and other ‘free gains’ are

often enough to fulfil the entire heating demand. According to Zhao & Magoulès (2012),

predicting energy consumption of buildings in general is being complicated because of the

variety of parameters that influence it. Regardless of the type of building, one can think of

ambient weather conditions, building characteristics (e.g. construction and thermal

insulation), lighting, HVAC and occupancy.

14
Ben-Nakhi & Mahmoud (2004) used neural networks to forecast next day cooling loads for

three public buildings in Kuwait. Their only input parameter was outdoor dry bulb

temperature. They had access to five year’s worth of data. 80% of the first four years was

randomly used for training, 20% for testing. The data on the last year was held out as

“production data”. The authors achieved excellent prediction accuracies with R2 values

exceeding 0,90. They conclude that NN’s are a relevant tool for cooling load prediction as

they require fewer input parameters than simulation models.

In the study of González & Zamarreño (2005) a two-phase neural network was built to

predict building energy load. The researchers used temperature parameters, time-related

parameters (to identify weekends and holidays) and load parameters. They found that the

load at time t was the most important predictor of the load at time t+1. Their neural network

utilised a hybrid algorithm, five neurons of the hyperbolic tangent type and 1000 training

cycles. The model performed very well. The authors noted that it appears that, for this type

of problems, many layers nor many neurons per layer are required.

Cho et al. (2004) mention three popular weather parameters for prediction purposes, namely

dry-bulb temperature, solar radiation and humidity. These authors measure heating energy

consumption between November 2000 and March 2001 for an office building in Daejon,

South-Korea and tested the influence of measurement length on prediction accuracy. Their

research shows the necessity of high quality input data that was gathered over as long a time

period as possible, as the outcomes indicated a 94 percentage point decrease in prediction

error between 1 day and 90 day measurements.

Interestingly, in the research of Datta et al. (2000) the best results were yielded using data

from ‘just’ the previous month. Training the network with more data (e.g. 4 months) resulted

in higher prediction errors. This somewhat contradicts the findings of Cho et al. (2004). The

most explaining variable proved to be time of the day, followed by interior and exterior

‘environmental conditions’ such as humidity and temperature. The best performing ANN

was able to realise half hourly predictions with an accuracy of 95%.

Datta et al. (2000) applied an ANN with one hidden layer to predict supermarket energy

consumption. The study focused on electricity consumption and is very much practice-

oriented. The authors state that energy reduction can be accomplished only through better

15
insight into both consumption patterns and energy demanding equipment in relation to

external parameters. The ANN model performed much better than both a linear and a

polynomial regression model. Also, according to Datta et al. (2000), ANN’s can achieve

higher levels of accuracy than simulation models.

Yuce et al. (2014) used ANN’s to predict energy consumption for an indoor swimming pool

in near real time. The goal was to achieve energy savings. The authors used an ANN as it is

fast, something that was required because the predictions were used in the Building

Management System (BMS). Their object of research is located in Rome, Italy and includes a

25/16/1.6-2.1m pool, a 16x4x1m pool as well as some other facilities. The building also

possesses a couple of energy generation systems such as a biomass plant, a diesel plant and

40 square metres of solar panels. Several subsystems of the building were taken into account,

including the air treatment subsystem, the water treatment subsystem and a building control

subsystem. No weather data were used. The dependent variables/objectives chosen by these

researchers were split into electricity consumption, thermal energy consumption and

thermal comfort (PMV). The researchers have determined and documented the ANN

settings and the best performing algorithms carefully, which is valuable for the present

research. The Levenberg-Marquadt backpropagation algorithm performed best. Further, the

authors yielded the best results using two hidden layers, 1000 training cycles and a learning
rate of 0,01.

Fan et al. (2014) developed models for the forecasting of both next-day energy consumption

as well as next-day peak power demand. They used a three-step approach consisting of

outlier detection, recursive feature elimination (RFE) and final model development. Most

detected outliers were related to public holidays. The RFE had a positive effect on both

computing times as well as model performance. Also, it prevented from overfitting. Using an

ensemble model constructed of eight popular prediction algorithms, the researchers were

able to reach a high level of accuracy with a MAPE of 2.32% and 2.85% for total consumption

and peak demand respectively. For all chosen variables, lags for 1 to 15 days were also used

as potential input variables. For the separate models, support vector regression and random

forests performed best, whilst linear regression and ARIMA performed poorly. The best

performing separate models also contributed most to the ensemble models. The paper stands

16
out because of the high number of meteorological variables that was incorporated into the

model.

Whilst most research focuses on a single type of objects (e.g. office buildings, supermarkets),

the study of Kalogirou et al. (1997) stands out because the researchers trained an ANN using

data of 250 different cases varying from ‘small toilets to large classroom halls’. The DV is the

heating load and the IV’s consist solely of room characteristics such as area and the presence

of windows. Given the variety of data the model is trained and tested with, the accuracy is

remarkably high (over 90%). Also, the authors note that the ANN allows for very fast

prediction times. The authors used different activation functions depending on the layers:

their ANN consisted of an eight-neuron input layer, a ten-neuron gaussian hidden layer, a

ten-neuron tanh hidden layer, a ten-neuron gaussian hidden layer and a 1 neuron logistic

output layer.

Newsham & Birt (2010) specifically researched the influence of occupancy on ARIMA energy

(in this case electricity) forecasts. They gathered this data using a variety of sources,

including wireless sensors, door sensors, motion sensors, carbon dioxide sensors and the

number of logins to the network. They conclude that, although accuracy slightly improves

when occupancy levels are taken into account, only the login data had explanatory power.

All other occupancy measures proved to be ineffective. The authors explain this mainly

through the somewhat unique nature of the building and the fact that a lot of data had to be
imputed.

Lai et al. (2008) used SVM’s to forecast electricity consumption for residential buildings. 15

months’ worth of data were used to train and test the model. Climate data was incorporated

into the model, as well as building related characteristics (e.g. building date, total surface and

overall heat transfer coefficient). The model performed well. The most important driver was

the outdoor temperature.

Taieb & Hyndman (2013) participated in a 2012 Kaggle load forecasting competition.

Although this competition entailed prediction of electricity load, there are many potential

methodological similarities to the present study. Their data came from a US utility provider

and entailed 20 zones. Also available was temperature data for 11 weather stations. The

relative locations of both zones and weather stations was unknown and needed to be

17
discovered. The authors incorporated the current temperature and lagged temperatures into

the model. They also used the minimum and the maximum temperatures in the current and

previous day as well as the average temperature for the current and previous day and for the

previous seven days. Lastly, they included the lagged values of the demand itself into the

model (with the same temporal lags as for temperature). In the end, the authors used a total

of 24 models (one for each hour) based on a component-wise gradient boosting technique

with the main predictors being current and past temperatures and past demand. For past

demand and temperatures, lag variables of at most -7 days were used. Using separate

models for every hour has been applied before successfully, as the authors note, by a.o.

Ramanathan et al. (1997) and Fay et al. (2003). The results clearly show the predictive power

of the most recent previous demand(s). Especially for the first three hours ahead, this was by

far the strongest predictor. During working hours, however, (lagged) temperature variables

proved dominant.

Fay et al. (2003) have dedicated their paper entirely to the aforementioned decision for the

hourly models: whether a comprehensive model for all hours of the day or a separate model

for each hour should be built. Their conclusion was that, except for four specific hours of the

day, the sequential approach (one model) outperformed the parallel approach (24 models).

Their final solution, a combination of the sequential and parallel model, performed even
better.

Fan et al. (2017) compared seven types of prediction techniques to predict 24 hour ahead

building cooling load. Multiple linear regression and elastic net functioned as performance

benchmarks. The focus of the paper is on deep learning methods. Unsupervised and

supervised deep learning methods were combined to enhance the prediction quality. The

object of research was a 11.000 m2 – largely air conditioned - school building. One year of

data for the year 2015 was available with 30 min intervals. Until this study, the potential of

applying deep learning methods to building energy load prediction was largely unknown.

Apart from the deep learning model performing very well, the paper stands out because of

the novel way in which feature selection was handled, as discussed in Section 2.2. Another

interesting characteristic of this study is the fact that the authors used an iterative approach to

predict the 24 hour ahead cooling loads. In essence this means that every one-hour

prediction is added to the 24 inputs used to predict the next one-hour prediction. Up to ten

18
hidden layers were used for developing the model. Adding more hidden layers does, as the

results show, not necessarily result in better performance. Specifically, the best results were

achieved using just two hidden layers. These authors explain this behaviour through data

scarcity: their one year dataset with 30 min intervals is likely not large enough to apply NN

models with such depth to. Both linear methods failed to perform very well. This can easily

be explained by the underlying relationships not being linear, which is why these methods

struggle. Overall, the “extreme gradient boosting” performed best with an CV-RMSE

(coefficient of variation of the RMSE, a way of comparing RMSE’s which in themselves are

scale-dependent) of 17,8% on one dataset. The computational load for the more complex

models is, however, much higher. Another interesting finding, in light of the feature

selection methods covered earlier, is the fact that the linear methods performed best on the

raw data, whilst the more advanced non-linear methods performed better on the high level,

nonlinear abstractions of the data generated through deep learning feature selection. Whilst

using engineering knowledge to select features can work better on one case than on another,

feature selection using deep learning will, these authors believe, almost always result in a

performance boost over using the raw feature set.

A great unknown to models such as the one we will develop, is air infiltration (the air

leakage into the building). This is problematic because it has a significant impact on the

building’s energy demand. A very thorough review of this topic has been written by Younes

et al. (2012). Some important conclusions in this paper are that air infiltration accounts for

25% to 50% of heating and cooling load in commercial buildings, and that for low-rise

buildings, the wind gusts at the building façade are an important driver of air infiltration,

whilst for high-rise buildings, the most important factor is the so-called stack effect. This

effect describes that for such buildings, during summer, the hot air rises to the top of the

building, thus creating a higher pressure in the top of the building and a lower pressure at

the bottom. As a result of this, air will infiltrate the building in the bottom area to correct the

pressure difference between the inside and outside air, and leak out of the building in the top

area.

19
2.4 Relevance
One of the most important benefits of accurate energy consumption predictions is the ability

to better align local energy generation with local energy demand. Accurately predicting

energy demand also allows for an economical gain through energy purchasing plans. This

holds at multiple levels, for instance on the level of the consuming party but also on the level

of the power provider. The power provider also benefits from this kind of research because

once predictions become better, grids can be dimensioned more accurately (less over-

dimensioning is required). A good example of the relevance of energy load

prediction/forecasting is given by Datta et al. (2000). These authors indicate that electricity is

purchased by most UK supermarkets on a half-hourly basis; a penalty is incurred in case in

case the prediction was off. This is a nice example of the monetary incentives of being able to

forecast energy consumption. The study also indicates that some load-shifting to off-peak
periods can be realised through accurate forecasts.

The aforementioned benefits, however, mainly hold for electricity load prediction. This

particular study has as its largest stakeholder DWA. The main insight they will extract from

it is whether data-driven models can perform at a similar level to physical models for this

specific type of problem (predicting heating and cooling loads for benchmark purposes).

More generally, however, these prediction models can ‘evaluate a proper operation and

audit retrofit actions’ (Neto & Fiorelli, 2008). As Kalogirou et al. (1997) point out, data driven

models can (as opposed to simulation models) be particularly valuable for countries where

accurate thermal properties of building materials tend to be unavailable.

Energy prediction models can be used for ‘fault detection and diagnosis, operation

optimizations and interactions between buildings and smart grid’, according to Fan et al.

(2014). Energy prediction models can function as a performance benchmark, enabling firms

like DWA to quickly respond to faulty behaviour or component failure. This is extremely

relevant as it is known that the energy performance of buildings degrades significantly over

time. First and foremost, this is caused by components wearing out, thus losing efficiency

and eventually failing altogether. However, decreasing energy performance is also often

caused by the building management and maintenance being hand-over to another party, that

is not entirely acquainted with the equipment in that specific building. (Wisse, 2017).

20
Although, as shown, data driven models have many benefits, there are also some drawbacks

to using them. One of the largest limitations is that the effects of retrofitting actions to the

building envelope cannot be accurately predicted. Using a physical model one can, with

relative ease, predict for instance what the energy savings of replacing all glass with triple-

isolated glass will be. The same estimation cannot be given upfront through data driven

models. Another limitation, that is very much applicable to the benchmark scenario for this

paper, is the fact that one can never reliably establish a baseline figure. In other words, when

the model is trained and the building at that point already malfunctions or contains

malfunctioning components, this behaviour is included into the model as being the

benchmark value. In this sense, a physical model remains valuable, as the performance with

all components functioning as they should can be predicted with relative accuracy. (Wisse,

2017).

21
3. Literature review II: ANN’s & ANN parameters
As ANN’s will form the backbone of our research, in this chapter, we shall discuss what
ANN’s are, how they work and which parameters are important in tuning them.

3.1 ANN’s
ANN’s exist in both hardware as well as software (algorithm) versions, with the former

being known as Hardware Neural Networks (HNN’s). We will be using an algorithmic ANN

for this research. A popular definition of a neural network, as given by Dr. Robert Hecht-
Nielsen, the inventor of one of the first neurocomputers is:

“… a computing system made up of a number of simple, highly interconnected

processing elements, which process information by their dynamic state response to

external inputs.”

(Caudill, 1989)

ANN’s are organized in layers. Each layer contains nodes that are connected to the nodes in

the adjacent layers, but not to the nodes in its own layer. Each of the nodes possesses an
activation function. A typical neural network is depicted in Figure 1.

22
Input Hidden Hidden Hidden Output
layer layer 1 layer 2 layer 3 layer

Figure 1: Typical ANN structure.

In the illustrated ANN, there is one input layer and one output layer and there are three

hidden layers. Through the input layer, the patterns are presented to the ANN. The

processing and model generation is being performed by the hidden layers. The lines between

the hidden layers are weighted connections. In the output layer, the final solution (in our

case, the estimation of the energy loads) is given. The weights of the connections between the

hidden layers are determined by the learning rule that the user gives the ANN. In the

weights, the model’s ‘knowledge’ is stored (Kalogirou et al., 1997). Its structural resemblance

of human brains as well as its ability to learn by itself are the reasons for the term ‘neural’.

In the present study, we apply a backpropagation neural network (BPNN), meaning that

errors are being backpropagated. Backpropagation algorithms are the most popular

algorithms used for ANN’s (Kalogirou et al., 1997). This type of network makes use of the

‘delta rule’: each learning cycle the activation flow of outputs goes forward and the

adjustment of weight goes backwards. In the first cycle, the ANN chooses the weights

randomly. In the second (and indeed, nth) cycle, it compares its output to the actual value

and adjusts the weights accordingly. (wisc.edu, 2017)

23
The activation function that exists at each node in the hidden and output layers is what gives

the ANN it’s non-linear abilities. In the ANN we use, the activation function is a sigmoid

function, resulting in the input of that node (that can theoretically range from -∞ to ∞) being

mapped to a value between 0 and 1:

1
𝛷(z) = 1+ 𝑒 −𝑧

According to Ben-Nakhi & Mahmoud (2004), this activation function is the most popular and

oftentimes the most effective. In more advanced deep learning models, the activation

function can be different for each layer (Gelston, 2017). This has been applied for instance by

Kalogirou et al. (1997).

After having been trained, the neural network can be applied to unseen data. At this point,

no more backpropagation of errors is required (this would also be impossible, because the

actual values are unknown) thus the network is set to only perform forward propagation.
The output now is the actual DV prediction.

Neural networks commonly have relatively few hidden layers. Deep learning models on the

other hand employ more hidden layers, assigning specific tasks to each layer (Fan et al.,

2017). Adding more hidden layers/neurons to the model does not necessarily improve

results. As Gelston (2017) describes, more hidden neurons will result in the ability of solving

more complicated problems. Too many hidden neurons, however, can result in overfitting or

decreased model performance. Finding the right number of hidden layers, then, is – like

many of the parameter tuning for an ANN – a process of trial and error.

BPNN’s have been criticized for being ‘the ultimate black boxes’, because a user has no

influence whatsoever on the model training, other than setting some initial parameters. Also,

depending on the number of training cycles, observations, input parameters and available
hardware, they can be quite slow. (wisc.edu, 2017)

3.2 ANN parameters


Tuning SVM’s and ANN’s is one of the most challenging parts of creating an accurate yet

generalizable model (Meyer, 2017). Two of the most important parameters of the ANN are

the number of hidden layers and the number of training cycles. In line with Fan et al. (2017),

24
we discovered that for this type of problem, using fewer hidden layers outperforms using

more of them. As these authors explain, the problem at hand is not so complex that adding

many hidden layers improves the results. In fact, our results became significantly worse,

once more than one single hidden layer was added: using the same dataset we experienced a

61% drop in accuracy with a second hidden layer and an additional 60% drop when adding a

third hidden layer.

Additionally, the results were influenced heavily by altering the number of training cycles.

Generally, the results improved by increasing this number. However, as Gelston (2017)

describes, overdoing this will specialize the model too much and restrict it from being able to

find more general patterns in unseen data. Although we apply cross-validation to decrease

overfitting risks, we did limit the number of training cycles to 1000. Compared to 500

training cycles, this still gave a significant boost in accuracy, kept the computation times

reasonable and would, we believe, not overfit the model to the known data. Also, it was in

line with the findings of González & Zamarreño (2005) and Yuce et al. (2014), who also used
1000 training cycles.

Other relevant ANN parameters are the learning rate, the momentum and decay. The

learning rate refers to the amount of change that is allowed to happen to the weights in each

training cycle. We achieved the best results using a learning rate of 0,3 with a momentum of

0,2. The momentum of a neural network refers to the addition ‘of a fraction of the previous

weight update to the current one’ (rapidminer.com, 2017). Its purpose can be explained as

follows: when too high a learning rate is used, the “best” solution can be underestimated in

cycle x only to be over-estimated in cycle x+1. What follows is an ‘oscillating pattern’ and no

convergence to the optimum (McCaffrey, 2017). The momentum parameter exists to prevent

this. By adding a fraction of the previous weight update, the ANN is able to ‘jump over’ such

local error minima, thus not getting trapped. By the time the network is close to the global

minimum (the optimal solution), the previous weight update will be very small so that the

ANN does not overshoot the global miminum in the same way it does with local minima

(McCaffrey, 2017). Applying decay – meaning that towards the final iterations, the learning

rate is being decreased – had a significant positive effect on our results.

25
4. Research objective and conceptual model
In this chapter, the research objective and research questions are being outlined, as well as
the conceptual model.

4.1 Research objective


The objective of the present study is to come up with a benchmark model for the heating and

cooling load of a Dutch office building that approximates physical reality with sufficient

accuracy.

4.2 Research questions


The question that we want to answer in this thesis is the following:

How can we come up with a model that accurately predicts daily/hourly

heating/cooling loads for a Dutch office building using external conditions?

The following sub-questions can be formulated:

(1) Which factors drive heating/cooling load for this building?

(2) Which types of models can be used to model this behaviour?

(3) Which model suits this type of problem best?

(4) What level of accuracy can be obtained?

Throughout this paper, the questions will be addressed. In the conclusion, we shall explicitly

answer the sub-questions, and thereby the main research question.

26
4.3. Conceptual model
Figure 2 depicts the conceptual model.

Figure 2: Conceptual model.

As the graphic shows, we have, based on the literature review and the interviews with

DWA, identified several predictors that we expect to be relevant for the model. We have split

them into four subclasses: building envelope, internal variable parameters, weather

parameters and temporal parameters.

The model we construct will be building specific, which means that (especially) the building

envelope will be ‘modelled’. The other categories (e.g. occupancy, weather, time) will be fed

to the model as input data, on which the heating/cooling load predictions will be based. The

temporal parameters in isolation have no effect, but will be included to account for

daily/weekly seasonality patterns that result from the other variables (e.g. occupancy,

working hours).

27
5. Case study
An overview of the building and its properties will be given in this chapter. Furthermore, the
two most important subsystems will be described in detail.

5.1 Building and subsystems


The object of research for this thesis is a medium size office building located in the

Netherlands. The gross floor area of the building is around 6.300 m2. The building is

comprised of a large hall, and a large office section behind that. It is designed to

accommodate about 210 people. The building is equipped with a variety of energy efficiency

features which, although very effective, complicate the model building. All data we use was

used by the monitoring software package Simaxx.

The illustration below, that was provided by DWA, shows a selection of climate control

systems that buildings can be equipped with (note that this building does not possess all the

‘options’) as well as the general system setup for this building.

Figure 3: General set-up and climate systems.

28
As Figure 3 shows, gas or electricity is drawn from the main grid/network, and processed to

heat or cold in a form of power plant, which is usually situated at building level (e.g. boilers,

chillers, heat pumps, aquifers), but sometimes centralised (city heating/cooling). This is then

transferred through the building by means of radiators, floor heating/cooling and Air

Handling Units (AHU’s). Other, more sophisticated techniques that DWA often uses in their

designs include concrete core activation, climate ceilings and active chilled beams.

In this paper, we will be focusing on the section indicated by the dashed blue line in Figure 3.

As a result, we will not be examining the heat/cold generation part, meaning that

inefficiencies in that part of the system will be outside of the scope of this document. Also, it

is repeated here that we will specifically be focusing on heating and cooling load, not on

electricity consumption.

In the case of our building, two specific subsystems will be analysed, namely the floor

heating and cooling and the AHU’s. Together, these, for this building, comprise the methods

of heating/cooling the building. We will, however, give most attention to the aggregate

circuit, depicted in Figure 3 as the horizontal red and blue arrows. It is also noteworthy that

for heating, there is, for some special rooms in the building, a small group of

radiators/convectors present, as an addition to the AHU’s and the floor heating system.

However, these systems only make a small contribution to total heating load (about 7% of

the total; Wisse (2017)).

In a physical sense, the heating and cooling demand parameter is exactly the same, except

for both being each other’s inverses. However, in an engineering sense, the cooling and

heating are delivered through two separate circuits. It is therefore possible for the building to

ask for heat and cold at the same moment. This would for instance happen when the AHU’s
ask for cold, but the floor heating system asks for heat at the same moment. (Wisse, 2017)

5.2 Air handling units: heat recovery system


The building possesses multiple heat recovery systems, but the most important one is a heat

recovery wheel. The function of this wheel is to transport heat from the return air flow to

heat the supply air flow (or vice versa). This thermal wheel, as it is also called, therefore is

29
located in the AHU circuit, not in the water circuit. To be able to transfer lots of heat energy,

the heat wheel consists of a very fine structure, in for instance a honeycomb pattern. The

wheel is being driven, slowly, by an electric motor to allow the heat to be transferred from

one flow to the other. The efficiency of the technique is rather good, compared to other air-

to-air heat exchangers. (Wisse, 2017)

5.3 Floor heating/cooling


The floor heating has been divided into two groups: hall and office. The reason for this is that

the central heating circuit is feeding 2 AHU’s as well as 2 floor heating and cooling groups.

One of these groups feeds the offices, the other group feeds the hall. Again, it is possible that

one of the groups asks for heat, whilst the other group asks for cold. (Wisse, 2017)

5.4 Residual electricity consumption


As mentioned above, we will not research the electrical part of the system. There is one

exception to that statement though: as we saw in the literature review, it has been widely

accepted that occupancy is one of the most important drivers of building heating/cooling

load (e.g. Yang & Becerik-Gerber (2017) and Fan et al. (2017)). This is the case because of

metabolism but also because of use of equipment, expelling heat (Yang & Becerik-Gerber,

2017). Also, interior lighting – which usually correlates with occupancy, for instance through

proximity sensors – has a significant influence on the heating/cooling required (Wisse, 2017).

While, as Fan et al. (2017) explain, actual occupancy figures are rarely available, they can be

imputed from other, known data, namely time variables. This is the case because occupancy

is usually very much correlated with time. Whilst we will be using time variables in our

models, we also have available two other indicators of occupancy, namely electricity

consumption and setpoints.

Although the measured electricity consumption is, in essence, an aggregate number,

comprising of many occupancy-dependent and occupancy-independent components, the

electricity consumed by the heat pumps – by far the largest occupancy-independent

component – is also measured. By subtracting this from the aggregate number, we expect to

30
get a rough approximation of occupancy. The largest limitation of using this data is the fact

that the power consumption of the AHU’s is also included in this number, and accounts for a

relatively large share of it. Since no data on AHU power consumption is available for this

building, we cannot correct for this and thus need to treat the residual electricity

consumption as a rough estimator of occupancy.

Through the setpoints/thermostats, we have available the number of rooms asking for

heating/cooling at any given moment. This too can be used as an occupancy approximation,

as it is correlated strongly with building occupancy.

Of course, using all three estimators (time variables, electricity consumption and setpoint

data) is likely to introduce multicollinearity into the model. We chose to use residual

electricity consumption in favour of setpoint data, as it was somewhat unclear how well the
setpoint data reflected occupancy in the non-office area of the building.

31
6. Data description
In this chapter, an impression of the data that was made available to us by DWA will be
presented. Also, we will discuss missing data in Section 6.1.

The data we received for the analyses was gathered over a one-year period (01-01-2016 to 31-

12-2016) in a medium sized office building located in The Netherlands. This location

differentiates the present study from most of the existing research because of the Dutch

climate. Whilst in other studies the research focused specifically on either heating or cooling

load, our dataset, as will be shown below, consists of a “hot” and a “cold” period,

complicating the analysis. For the building, DWA had a virtual physical model available.

Our objective was to approximate the performance of that model.

For the one-year period, data was collected with a one-hour interval. Since the year 2016 was

a leap year, and since one hour was not recorded at all, the dataset consisted of 8.783
datapoints.

The data was divided into several files, more specifically for the AHU’s, the floor heating

and cooling (again divided into an office part and a hall part), the outdoor and desired and

actual indoor temperatures and the central heating and cooling circuits. Our focus was on

the central circuit data: in it were both the dependent values we had to explain as well as the

factors that influenced it most notably. Additionally, data on the overall electricity

consumption and on the heat pump energy consumption was also used as described in

Section 5.3.

It is customary to express cooling demand using a negative value. This makes it easily and
immediately visible that a certain number or graph depicts cooling demand.

The graph in Figure 4 depicts the relationship between outdoor temperature and

heating/cooling load. The figure shows that outdoor temperature is heavily correlated with

both. Furthermore, the relationship between heating and outdoor temperature appears to be

more linear than the one between cooling and outdoor temperature: the cooling load

displays a “fork”. The upper group contains mostly the weekend observations, when the

AHU’s are inactive, and the only cooling power is being drawn through the floor systems.

The lower group consists mostly of weekdays, when both systems are often active. Because

of the less linear relationship between cooling demand and temperature (compared to
32
heating demand), we expect the heating load models to perform better. This is in line with

the hypotheses of Kalogirou et al. (1997), who note that cooling load estimation is much

harder than heating load estimation.

Daily cooling and heating load vs. outdoor


temperature
Heating load Cooling load

4000
Energy demand (kWh)

3000
2000
1000
0
-100 -50 -1000 0 50 100 150 200 250 300

-2000
-3000
Outdoor temperature (0,1 °C)

Figure 4: Daily cooling and heating loads plotted against outdoor temperature.

The tables below depict all the variables we had available in the data files. For the main data

file and the weather file, also shown are some basic descriptive statistics. We have only

included in this section the information on the data parts that (a) we used in the main

analysis or (b) were in other ways informative. Descriptives of the other available data are
added in Appendix A.

33
Variable Unit of measurement Min. Max. Avg.

Date/time 01-01-2016 0:00 31-12-2016 -

23:00

Temperature supply 1 °C 0 18,95 14,58

Pressure difference 1 kPa 0 250 85,76

Temperature return 1 °C 0 22,91 18,19

Setpoint temperature °C 0 16 14,43

supply

Setpoint pressure kPa 110 110 110

difference 1

Volume flow 1 m3/h 0 31,15 3,80

Pump 1 steering % 0 73,71 22,15

Pump 2 steering % 0 73,71 23.15

Temperature difference °C -9,8 0,15 -2,75

Energy demand kWh -148,02 4,04 -16,25

Table 1: Central cooling circuit variables.

The boxplots in Figure 5 display the average cooling load and the variation therein for every

day of the week. Notably, the median value for the cooling demand is highest on Tuesdays,

but the interquartile range and the lower whisker are highest on Wednesdays. The latter can

be explained through meetings taking place on Wednesdays in the building, leading to

occupancy until 20:00 PM instead of 18:00 PM. Only two values are marked as outliers,

which is expected as the number of observations is small. The cooling demand on Saturdays

and Sundays are not only the lowest, but also by far the most stable, as depicted by the

narrow interquartile ranges.

34
Figure 5: Boxplot of the daily cooling loads.

The hourly cooling boxplots in Figure 6 also show some interesting characteristics in the

data. First of all, a large increase in the cooling demand arises on hour 8, spanning between

7:00 AM and 8:00 AM. This is the moment the system is set to start conditioning the building.

The cooling demand continues to rise throughout the day and peaks between 15:00 PM and

16:00 PM. Thereafter it drops back steadily. A lot of outliers are present in the 7th, 19th and

20th hours. The first could originate from users who are in before 7:00 AM and manually start

demanding cooling. The latter two can be explained through the working days ending later

on the Wednesdays, as was mentioned earlier.

35
Figure 6: Boxplot of the hourly cooling loads.

Variable Unit of measurement Min. Max. Avg.

Date/time 01-01-2016 0:00 31-12-2016 -

23:00

Temperature supply 1 °C 19,2 49,26 29,67

Pressure difference 1 kPa 0 70,62 31,40

Temperature return 1 °C 0 34,92 24,27

Setpoint temperature °C 20 45 28,37

supply

Setpoint pressure kPa 70 70 70,00

difference 1

Volume flow 1 m3/h 0 20,51 2,83

Heating power delivered kWh 0 242,78 30,97

Pump 1 steering % 0 100 23,99

Pump 2 steering % 0 100 26,28

Temperature difference °C -32,2 37,61 3,89

Energy demand kWh 0 242,73 21,83

Table 2: Central heating circuit variables.

36
Some remarks now follow regarding the meaning of and relationships between selected

variables for the main heating and cooling data.

‘Temperature supply 1’ refers to the temperature of the water that flows towards the air

handling units and the floor heating/cooling system. ‘Pressure difference 1’ is an indication

for the system being in use. Since the temperature variable has a value regardless of the

system being in use, the pressure difference is needed to examine that. ‘Temperature return

1’ refers to the temperature of the water coming back from the building. This too is a single

value for the water coming back from the AHU’s as well as the floor heating/cooling system.

‘Setpoint temperature supply’ is the setpoint for the temperature of the supplied water,

whilst ‘Setpoint pressure difference’ is a fixed target value of 70 kPa. ‘Volume flow 1’ refers

to the flow rate of the water in the main circuit. When > 0,1 m3/h, it indicates heat/cold

demand being present. When below that, the system can be regarded as idle. ‘Heating power

delivered’ (for heating only) is the energy being delivered, as monitored by the system.

‘Pump 1/2 steering’ is the steering signal for the heat pumps. A value above 0 indicates

system activity. ‘Temperature difference’ refers to the difference between the supply and the

return water flow. For this variable as well, a threshold of > 0,1 °C can be treated as a real
heating/cooling demand being present. Finally, ‘Energy demand’ is our dependent variable.

The energy demand variable (our DV) was mathematically calculated as follows: ED = 1,16 *

temperature difference * volume flow 1. This formula can be physically explained because

1,16 is the product of multiplying the density of water with its heat capacity. For the heating

data, there was also a measured value. The measured values were very close to the

calculated values, but there were small deviations. We were instructed to use the calculated

values.

Figure 7 shows that the daily heating load median value peaks at Wednesdays. Interestingly,

the heating demand for the Saturdays and Sundays is far less stable than the cooling

demand, as shown by the bigger box plots. No outliers are present.

37
Figure 7: Boxplot of the daily heating loads.

The boxplots for the hourly heating load as displayed in Figure 8 show a steady increase

between 0:00 AM and 10:00 AM. After this, the demand drops until 15:00 PM until it starts to

rise again until 17:00 PM. After that, the heating demand decreases to a median value of zero

at midnight. A surprisingly high number of outliers is present between 20:00 PM and 01:00

AM. This indicates very unstable behaviour for the hourly values and will likely cause the

models to perform worse than the models for the more stable daily loads.

38
Figure 8: Boxplot of the hourly heating loads.

Variable Unit of measurement

Date/time

Setpoint secondary supply °C

heating

Setpoint secondary supply cooling °C

# of askers for heat

# of askers for cold

Primary return temperature °C

Secundary supply temperature °C

Supply valve steering signal 1 %

Supply valve steering signal 2 %

Cold demand Binary2

Heat demand Binary

Table 3: Hall floor heating/cooling variables.

2Although these variables are binary in nature, there are certain datapoints where they contain a decimal value.
This is the result of the variable depicting the average value over the measured hour.

39
Variable Unit of measurement

Date/time

Setpoint secondary supply °C

heating

Setpoint secondary supply cooling °C

# of askers for heat

# of askers for cold

Primary return temperature °C

Secundary supply temperature °C

Supply valve steering signal 1 %

Supply valve steering signal 2 %

Cold demand Binary

Heat demand Binary

Table 4: Office floor heating/cooling variables.

The data for the floor heating and cooling subsystems depict an additional layer of

‘complexity’ in the system, namely the fact that the subsystem works in a democratic

manner. In other words: it cannot deliver heating to certain users and cooling to other users

at the same time. It therefore incorporates a ‘voting system’ that decides what will be

supplied, based on the number of askers for heating and cooling, respectively. This is also

the reason for the ‘Cold demand’ and ‘Heat demand’ being binary values: 1 means that this

type of conditioning has, for the moment, won. Having this data on individual rooms (e.g.

setpoint temperatures, number of askers), however, is valuable for the research. As Zhao &

Magoulès (2012) point out, one of the largest problems with this type of analyses has

traditionally been the lack of data on subsections of buildings. Having it present, thus, can be

beneficial.

In addition to the data provided by DWA, we also used weather data from the Dutch

national weather forecasting institute, the KNMI. There were two reasons for this:

(1) The DWA data only provided us with outdoor temperature measurements, whilst the

literature review showed that other variables such as solar radiation, wind speed and

humidity could also influence energy demand significantly, and;

40
(2) The recorded temperatures were not necessarily perfectly accurate; KNMI data

would probably be more reliable.

We retrieved the 2016 hourly and daily weather data datasets from the KNMI website for the

station nearest to the building. Table 5 below depicts a summary of the variables that were

available. The yearly weather data dataset was even more extensive, but did not contain

additional useful features.

Variable Unit of measurement Min. Max. Avg.

Date(/time) 20160101 20161231 -

Avg. wind direction over last 10 ° (990 for variable) 0 990 194,38

minutes of past hour

Avg. wind speed over past hour 0,1 m/s 0 130 34,00

Avg. wind speed over last 10 minutes 0,1 m/s 0 140 34,11

of past hour

Max. wind gust over past hour 0,1 m/s 0 250 61,00

Temperature at 1,50 m 0,1 °C -101 349 108,17

Min. temp. at 10 cm in past 6 hrs 0,1 °C -123 298 76,38

Dew point temp. at 1,50 m 0,1 °C -107 220 73,31

Sunshine duration over past hour 0,1 hour 0 10 2,06

Global solar radiation over past hour J/cm2 0 326 43,55

Precipitation duration over past hour 0,1 hour 0 10 0,81

Hourly precipitation amount 0.1 mm (-1 for <0.05 mm) -1 181 0,88

Air pressure at mean sea level 0,1 hPa 9.790 10.456 10162,06

Horizontal visibility [categorical] 0 83 61,76

Cloud cover Octants (9 for sky invisible) 0 9 5,78

Relative athmospheric humidity at 1,50 % 23 100 80,97

Present weather code [categorical] 1 92 29,31

Fog Binary 0 1 0,05

Rain Binary 0 1 0,21

Snow Binary 0 1 0,00

Thunder Binary 0 1 0,00

Ice formation Binary 0 1 0,02

Table 5: Weather data summary.

41
Although being the best data available, significant deviations are still expected, given the

distance of ~30 km between the KNMI station and the building. Particularly, we expect the

deviations for the solar radiation figures to have a large impact on the accuracy of the hourly

predictions (Wisse, 2017).

6.1 Missing values


Missing values were, at least in the main dataset, a problem: of the 8.783 values that should

have been recorded 2.170 had not been recorded due to the measurement software failing.

This left us with only 6.613 useful hourly observations. Furthermore, the missing data was

mostly concentrated in two blocks. This meant that imputation would not be possible,

especially considering the fact that we only had data for one year available. For these

reasons, imputing values would not be statistically valid.

The missing hourly data affected 124 days, meaning we had 242 daily observations available.

The ‘gaps’ in the data are clearly visible in the graphs below.

When we look at the daily central cooling load graph depicted in Figure 9, we see that the

cooling load is, of course, significantly higher during the summer months. There is, however,

a demand for cooling throughout the year.

Central cooling circuit - daily


0
12-okt
27-okt
12-sep
27-sep

11-dec
26-dec
11-nov
1-jan
16-jan
31-jan

1-mrt
16-mrt
31-mrt

26-nov
15-feb

15-apr
30-apr
15-mei
30-mei

14-jul
29-jul
13-aug
28-aug
14-jun
29-jun

-500
Energy demand (kWh)

-1000

-1500

-2000

-2500

Figure 9: Daily data for the central cooling circuit.

42
The same cannot be said for the heating demand, which is depicted in Figure 10. During the

summer months, roughly June until September, there is no heating demand for most of the

time. Also, again as expected, the heating demand increases sharply during the winter

period.

Central heating circuit - daily


3000

2500
Energy demand (kWh)

2000

1500

1000

500

0
12-okt
27-okt
12-sep
27-sep

11-dec
26-dec
1-jan
16-jan
31-jan

1-mrt
16-mrt
31-mrt

11-nov
26-nov
15-feb

15-apr
30-apr
15-mei
30-mei

29-jun
14-jul
29-jul
13-aug
28-aug
14-jun

Figure 10: Daily data for the central heating circuit.

Within the weather dataset, not all parameters had been recorded for every hour, but for the

variables we needed, there was a valid observation available for all 8.783 hours. Figure 11

below depicts what we expect to be the most important weather variable for predicting the
cooling and heating demand: mean outdoor temperature.

43
Mean outdoor temperature - daily
300

250

200
Temperature (0,1 °C)

150

100

50

11-dec
26-dec
15-mei
30-mei

12-okt
27-okt
12-sep
27-sep
1-mrt
16-mrt
31-mrt

11-nov
26-nov
1-jan
16-jan
31-jan
15-feb

13-aug
15-apr
30-apr

14-jun
29-jun
14-jul
29-jul

28-aug
-50

-100

Figure 11: Daily outdoor temperature data.

The two variables that are most at risk of being collinear are temperature and solar radiation.

However, solar radiation was expected to be an important driver for the heating and cooling

loads (Wisse, 2017). The graph below shows that although the trend is definitely collinear,

the behaviour is dissimilar enough to allow both variables into the model. Also, the variables

have been used alongside each other in numerous previous studies (e.g. Neto & Fiorelli,

2008; Cho et al., 2004).

Mean outdoor temperature vs. global radiation -


daily
300 3500
250 3000
Global radiation (J/cm2)
Temperature (0,1 °C)

200
2500
150
2000
100
1500
50
1000
0
16-nov
15-okt
31-okt
13-sep
29-sep
2-feb

5-mrt
21-mrt

2-dec
18-dec
1-jan
17-jan

18-feb

28-aug
8-mei

12-aug
6-apr
22-apr

24-mei
9-jun
25-jun
11-jul
27-jul

-50 500

-100 0

Mean temperature Global radiation

Figure 12: Daily outdoor temperature plotted against global solar radiation.

44
7. Methodology
In this chapter, we will discuss the methodology that has been used to create the models.

Figure 13 depicts a schematic overview of the process. The steps will be elaborated upon in

the following sections.

Figure 13: Schematic overview of methodology.

7.1 Data preparation


Before starting building the models, we made some modifications to the data. First of all, as

it was dispersed over many files (e.g. year quarters, subsystems) we integrated everything

into one file. Next, we manually inserted an (empty) observation for January 1, 2016, 02:00, as

this was the only value that was missing. It would have caused problems of not all days

contained 24 observations. We combined the hourly data with the hourly KNMI dataset.

Further, we removed the differences between the DWA data and the KNMI data with

regards to handling the Daylight Saving Time (DST). Using pivot tables, we converted the

hourly data (which was the only data available in case of the DWA datasets) to daily data.

This daily data was then combined with the daily KNMI dataset. Lastly, we altered the date

variables to a more generic format and extracted the weekday variable (Mon-Sun) for all data

as well as the hourcode variable (1-24) for the hourly data. We did not remove any outliers as

45
there was no valid reason to assume that these observations had been caused by system

malfunctioning. Also, as was discussed in Section 6.1, no imputations were made for missing

values, because (a) the missing data mostly occurred in large blocks and (b) we only had data

for one year. These factors combined made meaningful imputation impossible.

7.2 Seasonality
The data analysis in Chapter 6 revealed a strong weekly seasonality: The average energy

demand for cooling and heating would peak on Tuesday and Wednesday, respectively, and

be lowest on Saturday and Sunday. The plot also displayed an obvious yearly pattern:

cooling demand would peak in the summer, specifically somewhere in July and heating

demand would peak in the winter months.

For the hourly data, the heating demand would on average be highest between 8:00 AM and

9:00 AM. The cooling demand, on average, was highest in the afternoon, around 16:00 PM.

Since the outside temperature and the solar radiation figures peaked before that time

(around 14:00 PM), this already indicated inertia to be present in the building.

Taieb & Hyndman (2013) state that it is important to consider lagged temperatures in a

demand forecasting model due to thermal inertia in buildings. These researchers further

applied a log transformation to the demand, to stabilise the variance over time. Their results

show that boosting techniques perform very well for energy explanation/prediction models,

particularly because they appear to be relatively resistant to overfitting. Inspired by these

authors, we also took into account lagged variables for all our DV’s. The temporal lags were

chosen based on the literature review (e.g. Fan et al. (2014) and Taieb & Hyndman (2013)).

7.3 Feature selection


In this paper, we used engineering knowledge as the main instrument for feature selection.

We did not feed the models with all available features, because many of the parameters

physically could not have a relationship with the dependent variable, and including all

would make the models prone to overfitting. We therefore hypothesized a selection of

potentially relevant parameters based on the literature review and the interviews with DWA.

46
We did not choose for the deep learning methods used by Fan et al. (2017) because the higher

level, more abstract feature classes these would generate would introduce yet another ‘black

box’ and complicate deploying the model for benchmark purposes using a certain input

scenario.

In first instance, we selected the following features for our models:

Daily models Hourly models

Avg. wind speed Avg. wind speed

Avg. temperature Temperature @ 1,50 m @ tobs

Global radiation Global radiation

Air pressure Air pressure

Rel. humidity Rel. humidity

Residual electricity consumption3 Residual electricity consumption

Weekday Weekday

Hourcode

Table 6: Initial features.

Air pressure, although (rarely) used in other papers, was excluded afterwards, because its

influence could not easily be physically explained (Wisse, 2017) and because it had an

adverse effect on the performance of the models. Also, for the daily models, a variable was

added, namely maximum temperature. This had a significant positive effect on model

performance. Initially, cloud cover was also taken into account as a variable. Since this is, in

the KNMI dataset, an ordinal variable (expressed in octants, 1-8) dummy variables were

created. All of the models however performed slightly worse than they did without it, and

there was a lot of collinearity with the solar radiation parameter that was also present.

As we learned from the study by Younes et al. (2012), air infiltration is an important

determinant of heating/cooling load and thus needs to be taken into consideration. For the

building presently studied – which is a low-rise building – there is no “measurement” of air

infiltration. To account for this, we included wind speed into the model. The aforementioned

3No actual occupancy data is available, thus an approximation (residual electricity consumption) was
used. This is described in more detail in Section 5.4.

47
study showed that wind gusts at the façade are the main cause of air infiltration for low-rise

buildings.

The following features were selected to be included in the final models:

Daily models Hourly models

Avg. wind speed Avg. wind speed

Avg. temperature Temperature @ 1,50 m @ tobs

Max. temperature Global radiation

Global radiation Rel. humidity

Rel. humidity Residual electricity consumption

Residual electricity consumption Weekday

Weekday Hourcode

Table 7: Selected features for daily/hourly models.

Weekday and hourcode were primarily chosen as an approximation of occupancy, and to

incorporate the clear weekly seasonality into the model. Residual electricity consumption

(total building electricity consumption minus electricity consumed by the heatpump) was

chosen as a more specific measure of occupancy because, as discussed, it would fluctuate

with the level of occupation inside the building (e.g. through lighting and devices). Analysis

did not show too much collinearity with the weekday/hourcode variables, so we could safely

use all three.

Initially, we also included lagged variables of the heating/cooling demand itself into the
model. Literature (e.g. Taieb & Hyndman, 2013) showed that this would likely be a good
predictor of current demand. However, we removed these lags for two reasons: first, they
completely dominated all other variables. The predicted energy demand became so
dependent on the past energy demands, that the external variables (weather, occupancy,
time) became almost redundant. This also introduced large prediction errors during
transition periods. Second, after discussing this tactic with DWA, we decided it was
preferred to feed the model with only external variables. This way, a certain real-world
scenario could easily be given to the model as input in roughly the same way as for the
physical model. DWA decided that this would significantly increase value. This also makes
intuitive sense as a rising energy demand due to component wear could simply become part
of the baseline through the lags, invalidating the model for benchmark purposes.

48
7.4 Main models
We chose to build two types of models. The first type was a linear regression model. This

functioned as the benchmark model. The second type was a model based on an artificial

neural network (ANN). Apart from the fact that they suit the problem, we had a couple of

reasons for choosing an ANN approach over an SVM approach. First, according to Zhao &

Magoulès (2012) SVM’s can suffer from slow running speeds, which could be a problem if

the analysis would have to be executed regularly (for instance, on the level of a building

management system). As Yuce et al. (2014) point out, ANN’s are particularly suitable for this

kind of research as they can handle noisy and incomplete data. Our data had a lot of missing

observations. ANN’s do appear to have a high risk of overfitting when they are trained too

long – even more so than SVM’s or more traditional statistical methods. This was monitored

by always testing on a holdout set and by keeping the number of training cycles modest.

7.4.1 Linear regression model


To provide a baseline, we built a linear regression model to explain the drivers behind the

energy demand. This meant that we built four separate regression models, namely two daily

ones (heating and cooling) and two hourly ones. All models were built using the RapidMiner

7.5 software.

In the model, dummy variables were generated for the nominal variables in the data, namely

weekday in case of the daily data and weekday and hourcode in case of the hourly data. As

is customary, one of the dummy variables was removed, as it would be fully explained

through (the absence of) the others. Next, lag variables were created. The tables below depict

the lags we used for the daily and hourly data, respectively.

Daily models Hourly models

Avg. wind speed (1, 2 days) Avg. wind speed (1, 2, 4, 24 hours)

Avg. temperature (1, 2 days) Temperature @ 1,50 m @ tobs (1, 2, 4, 24 hours)

Max. temperature (1, 2 days) Global radiation (1, 2, 4, 24 hours)

Global radiation (1, 2 days) Rel. humidity (1, 2, 4, 24 hours)

Rel. humidity (1, 2 days) Residual electricity consumption (1, 2, 4, 24 hours)

Residual electricity consumption (1, 2 days)

Table 8: Lag variables for daily and hourly data.

49
The resulting feature set was filtered to exclude N/A examples. 10-fold cross-validation was

applied to (a) get the maximum out of the limited data available (especially daily) and (b)

remove the possibility of getting an overly optimistic or pessimistic outcome for the

performance measures through a “lucky” training/testing split. The folds were split using

stratified sampling. The linear regression model was configured to, itself, use the M5_prime

algorithm for further feature selection, and to eliminate collinear features with a minimal

tolerance of 0.05.

7.4.2 Neural Network model


To make as fair a comparison as possible to our benchmark model, we used the same

features and largely the same preparation steps for the neural network. As described in

Section 3.2, the neural network was set-up to use 1000 training cycles with a learning rate

(indicating how much the weights may change at each step) of 0.3 and a momentum (a

fraction of the previous weight update added to the current one, to prevent local minima) of

0.2. A huge improvement to the performance turned out to be given through enabling decay,

meaning that the learning rate is being decreased during learning. Lastly, shuffling (of the

input data, required for ordered data) was used and the input data was being normalized,

which is necessary since a value between -1 and 1 is required for the sigmoid activation
function.

To make the ANN’s more tangible, the visual representation and the description of the daily
cooling neural network are made available in Appendix B.

7.5 Post-processing: ‘Zero-inflation’


Already in early stages, it became apparent that the models did not handle the large span of

heating demand absence in the summer well.

To give an impression: for the heating/hourly dataset, 3.430 datapoints had the value zero -

more than half of the total datapoints available. To deal with this, we decided to try a two-

stage approach for the (stable) daily heating model. The first stage would be the actual

(regression/neural network) model, whilst afterwards, all values estimated below 250 would

50
be set to zero by the model. This threshold was chosen in an arbitrary manner, mainly

supported by the graphs.

Whether this second stage would improve prediction accuracy was uncertain. Because the

estimated values without such an approach are usually close to zero, it might not result in a

significant improvement of the model (Wisse, 2017). Also, the second stage obviously

introduces wrong corrections itself, potentially with larger impact than the former, minor

estimation errors. Nonetheless, we wanted to find out whether this approach could impact

the model positively and did so using both the benchmark linear regression heating model as

well as the neural network heating model.

7.6 Performance measures


We used three measures of performance. First of all, the Root Mean Square Error (RMSE)

was used as an expression of the average absolute error the model makes. This is an

interesting statistic, however, it has the issue of being scale dependent: in this case, it is

depicted in kWh. To overcome this, two other – scale-independent – statistics were also

considered. The R2 value indicates how much of the total variance in the DV is being

explained by the model. The CV-RMSE is an extension of the RMSE, commonly used in the

field of building energy prediction. It stands for coefficient of variation of the RMSE and is

calculated as follows:

𝑅𝑀𝑆𝐸
𝐶𝑉𝑅𝑀𝑆𝐸 =
𝐸̅

Where 𝐸̅ is the average actual heating/cooling load.

51
8. Results
In this chapter, we will discuss the results that were achieved using the regression models

and the neural networks. For each result, we will differentiate between excluding and

including the occupancy-related parameters (weekday, residual electricity consumption,

hourcode in case of hourly data) to examine the impact of adding these parameters.

The results we achieved with the final models, but without the occupancy-related

parameters, are summarised in Table 9.

Dataset Method RMSE R2 CV-RMSE

Daily cooling LR 306.907 0.642 0.594

Daily cooling NN 288.578 0.678 0.558

Daily heating LR 265.579 0.875 0.367

Daily heating NN 174.451 0.944 0.241

Hourly cooling LR 20.941 0.512 0.980

Hourly cooling NN 17.958 0.641 0.840

Hourly heating LR 36.304 0.411 1.179

Hourly heating NN 33.083 0.510 1.074

Table 9: Results without occupancy-related parameters.

Several observations can be made. First, in all cases (daily/hourly, heating/cooling,

RMSE/R2/CV-RMSE) the neural network models outperform the linear regression models.

Second, without the occupancy parameters, all models struggle to achieve acceptable levels

of accuracy, with the only exception being the daily heating neural network model, which

comes reasonably close to the ASHRAE recommended threshold (which is discussed in

Chapter 9) of 20%. Third, as was hypothesised earlier, the daily heating demand is much

easier explained through the weather parameters. The models struggle to achieve a decent

result for the cooling variables. The hourly demand as well is very difficult to predict with

some accuracy using just the weather parameters. Last, it is obvious that the hourly models,
especially for heating, perform poorly.

The results with the complete, final models (thus including occupancy parameters) are

summarized in Table 10.

52
Dataset Method RMSE R2 CV-RMSE

Daily cooling LR 212.529 0.829 0.390

Daily cooling NN 103.966 0.958 0.191

Daily heating LR 238.000 0.898 0.341

Daily heating NN 141.595 0.962 0.203

Hourly cooling LR 18.075 0.644 0.830

Hourly cooling NN 10.976 0.868 0.504

Hourly heating LR 29.027 0.626 0.938

Hourly heating NN 23.242 0.760 0.751

Table 10: Final results, including occupancy-related parameters.

Again, these statistics reveal some interesting properties of the models. First, it is obvious

that the occupancy parameters add explanatory power to the model, as all models perform

better than their counterparts without occupancy parameters. Second, even with the

occupancy parameters included, the hourly models fail to achieve decent levels of accuracy.

Lastly, the cooling models appear to benefit more from adding the occupancy parameters

than the heating models.

The graphs below visualize the performance of the daily regression models and the daily
neural network models.

Results regression with cross-validation - cooling


- daily
prediction(Sum of ED) Sum of ED

1000

500

0
1

127
10
19
28
37
46
55
64
73
82
91
100
109
118

136
145
154
163
172
181
190
199
208
217
226
235

-500

-1000

-1500

-2000

-2500

Figure 14: Results of the daily cooling regression model.

53
Results regression with cross-validation - heating
- daily
prediction(Sum of ED) Sum of ED

3000
2500
2000
1500
1000
500
0
1

127
10
19
28
37
46
55
64
73
82
91
100
109
118

136
145
154
163
172
181
190
199
208
217
226
235
-500
-1000

Figure 15: Results of the daily heating regression model.

As the above graphs show, the regression models perform reasonably well in situations

where heating/cooling demand is substantial, but fail to do so when loads are small or zero.

In the cooling model, a significant exaggeration for zero-load values is visible, whilst also in
the heating model, it is apparent that the model struggles to predict low demands.

Results NN with cross-validation - cooling - daily


prediction(Sum of ED) Sum of ED

500

0
1

127
10
19
28
37
46
55
64
73
82
91
100
109
118

136
145
154
163
172
181
190
199
208
217
226
235

-500

-1000

-1500

-2000

-2500

Figure 16: Results of the daily cooling neural network model.

54
Results NN with cross-validation - heating - daily
prediction(Sum of ED) Sum of ED

3000

2500

2000

1500

1000

500

217
1
10
19
28
37
46
55
64
73
82
91
100
109
118
127
136
145
154
163
172
181
190
199
208

226
235
-500

Figure 17: Results of the daily heating neural network model.

The ANN models, however, perform much better at this. For instance, the predicted values

for the heating demand are far closer to zero. Apart from the prediction errors in the low-

demand regions, the overall prediction accuracy is also significantly better.

Figures 18 and 19 visualise the performance of the hourly models. We have randomly

selected 2% of the dataset (~130 hours), as displaying all 6496 available datapoints in a

meaningful manner was not feasible within a single chart.

55
Results NN with cross-validation - cooling - hourly (partial)
prediction(Energy demand) Energy demand

20

0
1 5 9 13 17 21 25 29 33 37 41 45 49 53 57 61 65 69 73 77 81 85 89 93 97 101105109113
-20

-40

-60

-80

-100

-120

-140

-160

Figure 18: Partial results of the hourly cooling neural network model.

Results NN with cross-validation - heating - hourly (partial)


prediction(Energy demand) Energy demand

250

200

150

100

50

0
1
4
7
10
13
16
19
22
25
28
31
34
37
40
43
46
49
52
55
58
61
64
67
70
73
76
79
82
85
88
91
94
97
100
103
106

-50

Figure 19: Partial results of the hourly heating neural network model.

56
8.1 Zero-inflation
As discussed, due to almost no heating demand being present during the summer, a

significant part of the errors for the heating model appeared to originate from the wrong

zero-predictions. To improve this, we applied the zero-inflation tactic discussed in Section


7.5 to the daily models. The results are summarised below:

Dataset Method RMSE R2 CV-RMSE

Daily heating NN 140.701 0.993 0.094863

Table 11: Results of the zero-inflated daily heating neural network model

Results NN with cross-validation and zero-


inflation - heating - daily
Prediction(Sum of ED) Sum of ED

3000

2500

2000

1500

1000

500

0
217
1
10
19
28
37
46
55
64
73
82
91
100
109
118
127
136
145
154
163
172
181
190
199
208

226
235

Figure 20: Results of the zero-inflated daily heating neural network model.

The performance improvement is substantial. As the graph shows, almost all the errors of

the model around the zero-line have disappeared, whilst almost no new errors have been

introduced. The CV-RMSE value being below 10 percent indicates at the fit being excellent.

Overall, we can conclude that the heating model performs better than the cooling model. We

expected this to be the case as the cooling demand is less linearly related to the driving

variables such as temperature.

57
8.2 Additional validation: 2015 data
After the models had been built, we received from DWA data on the last 45 days of 2015.

Similar to Ben-Nakhi & Mahmoud (2004), we wanted to use this as “production data” since it

was completely new to the neural network. Although for the most of 2015 the control

strategies for the AHU’s had been different – invalidating the data for this purpose – the

control strategies from November 16 onward were similar to the 2016 setup. As a result, we

were able to validate the models on an entirely independent test set. We focused on the

neural networks and on the daily models. The performance we achieved using this test set is

summarized in Table 12, and visualized in Figures 13 and 14.

Dataset Method RMSE R2 CV-RMSE

Daily cooling NN 94.817 0.626 0.476

Daily heating NN 185.973 0.759 0.177

Table 12: Neural network performance on the 2015 validation set.

Results validation set (nov/dec 2015) - NN -


cooling - daily
100

0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39
-100

-200

-300

-400

-500

-600

Sum of ED prediction(Sum of ED)

Figure 13: Results of the neural network on the 2015 validation set (cooling).

58
Results validation set (nov/dec 2015) - NN -
heating - daily
2500

2000

1500

1000

500

0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39

Sum of ED prediction(Sum of ED)

Figure 14: Results of the neural network on the 2015 validation set (heating).

These results show that the heating model performs very well, but the cooling model

struggles. Note that although the absolute error is lower for the cooling model, this is caused

by the cooling demand itself being much lower. This behaviour can be explained through the

2015 test set consisting entirely of winter months.

8.3 Computational load


The models should at some point be able to run as part of a Building Management System

(BMS), in order for which they have to be computationally efficient. Although this was not

the primary focus of this study, we did include below the computation times for each of the

final models (including occupancy-related parameters) in Table 15. All data indicate training

and testing times combined. Testing times are normally a small component of total time.

Also, complete retraining of the model for every prediction does not need to happen. The

models were built using an Intel quad-core processor running at 3,60 GHz using 4 GB of

memory.

59
Model Computation time

Regression – heating – daily 0m4s

Regression – cooling – daily 0m4s

Regression – heating – hourly 1m31s

Regression – cooling - hourly 1m40s

Neural Net – heating - daily 0m22s

Neural Net – cooling - daily 0m30s

Neural Net – heating - hourly 48m59s

Neural Net – cooling - hourly 50m22s

Table 15: Computation times per model.

60
9. Discussion
The ASHRAE (American Society of Heating, Refrigerating and Air-Conditioning Engineers)

which is an American organisation that leads in developing standards and guidelines for

(among others) HVAC systems states in ASHRAE Guideline 14-2002 that the CV-RMSE for

the baseline models needs to be below 20% for daily models when < 12 months of data are

available (Bing, 2005) and below 30% for hourly models (Yin et al., 2010). These guidelines

are very strict: According to Kalogirou et al. (1997), accuracies above 90% are easily within

the acceptable bounds for design engineering purposes for these sorts of problems. As the

results depict, the linear regression models fail to achieve any decent levels of accuracy for

all four datasets. We expected this to be the case because the linear regression models are

unable to model the complex underlying interactions between the input factors. Also, from

the data analysis and the literature review we already learned that many relationships in this
context appear to be non-linear.

The neural networks performed far better and the daily models achieved a level of accuracy

that is compatible with the ASHRAE guidelines. The hourly models were unable to do so. As

we saw in Chapter 6, the hourly values are far more volatile than the daily values. The neural

network models gave an improvement over the regression models, but also did not reach the
ASHRAE recommended treshold of 30% for hourly models.

Including the occupancy-related parameters appears to have a greater positive effect on

cooling load prediction accuracy than on heating load prediction accuracy. This may indicate

that the heating demand results more from a fixed schedule, whilst cooling is more often

explicitly demanded by the users of the building.

Applying the models that were trained (using cross-validation) with the 2016 data to the data

for the last 45 days of 2015 lead to some interesting observations. First, the heating model

performed far better than the cooling model. This can be explained by these months being

winter months only, where heating is more predictable than cooling, based on external

parameters. We believe the results would have been reversed, had the holdout set been for a

summer period.

61
The performance of the neural network will only improve once it learns further as a result of

more data becoming available. Frequent network retraining will be necessary to benefit from

this.

Returning to our original research objective and questions, we can present the following.

A variety of factors drives heating and cooling load for buildings. The most important

weather-related variables are likely average and maximum ambient temperature, solar

radiation, wind speed and humidity. Also, occupancy and building envelope are important

factors. For the building researched in this study, the weather parameters as well as the

occupancy-related parameters proved to have significant explanatory power.

To model heating and cooling load for commercial buildings, the most popular data-driven

solutions are ANN’s and SVM’s, although other types of models have also been used

frequently and successfully. We have not been able to compare all relevant types of models

in this study, but we can confirm that linear regression techniques are unable to model the
complex, non-linear relationships that underlie building energy load prediction problems.

In terms of accuracy, the daily models we built outperformed the physical principles model.

Depending on the model, accuracies in excess of 95% can be obtained.

62
10. Theoretical and managerial contribution
This paper contributes to an already rich field of research in several ways. First, it is

relatively unique because of its object of research. Not many heating/cooling load prediction

studies have been performed in the Netherlands, or even, surprisingly, in mixed climate

environments. We believe we have made a solid starting point, although we have

suggestions for improvement (as will be discussed in the next chapter). The other

characteristic that distinguishes the present study from most existing literature is its

objective. The majority of existing research that involves ANN’s uses prediction and

forecasting models to accomplish savings through for instance purchasing contracts or load

shifting. The primary goal of this study, on the other hand, was to come up with a

benchmark model that can be ‘fed’ with a certain real-world scenario and will give a baseline

figure that can be used to for instance detect component malfunctioning.

From a societal perspective, research in this field can result in savings in energy consumption

and thus in CO2-emissions. As we discussed, a huge share of total energy consumption

originates from (commercial) buildings, making it a very relevant sector to achieve

significant savings through optimising the buildings’ energy performance.

From a firm perspective also, there are benefits from this and similar studies. First, it is

obviously in every firm’s best interest to keep its building energy consumption to a

minimum, without sacrificing comfort. An ‘easy’ way of doing this is carefully monitoring

the energy performance of the building. Our data-driven models show that a benchmark

model for this purpose can be set-up even without in-depth engineering knowledge and

with relatively little resources.

Second, firms like DWA could benefit from this research. Such firms often need to adhere to

energy performance contracts, meaning that a certain energy performance of the building is

‘guaranteed’ by the HVAC consultancy firm. Currently, energy performance is being

monitored and compared to a physical principles benchmark model to detect deviations.

Our models appear to be able to do the same job, but are far less time-consuming to build.

Also, they are more accurate and they can be re-trained regularly to increase performance

further.

63
11. Limitations and suggestions for further research
Although in this thesis we were able to create a model that is sufficiently accurate for daily

predictions of both heating as well as cooling loads, some limitations exist and we have

suggestions regarding potential future research. Both will be presented in this chapter.

The first limitation that needs to be acknowledged concerns the data that we had available.

Although a years’ worth of data is not necessarily too little, the large periods for which the

measurement system failed to record measurements does pose a problem. After all, it means

that for these periods (spanning multiple months) the model cannot be trained in an accurate

manner. Another clear limitation, that we discussed in Section 2.4, concerns data driven

benchmark models in general: faulty behaviour can already be included in the ‘baseline’. In

other words, for buildings as large as the one in this study, there is no way of knowing for

certain that all components work as they should at the time the data is being gathered. Even

a new building can already have some degree of malfunctioning present. A way of dealing

with this is by having a physical principles model alongside the data driven model. Of

course, this somewhat defies the primary purpose of the data driven model in the first place,

which is saving resources. However, given that we were able to outperform the physical

model in certain aspects, having both models as complements could still be a viable solution.

11.1 Suggestions for further research


For further research, we have the following suggestions.

First, looking at the results achieved on the 2015 holdout set, we believe that separate

summer and winter models would benefit accuracy greatly, because cooling and heating

behave completely differently depending on this condition. As such, the cooling models

struggle during the winter period, and the heating models during the summer periods.

Second, it can be interesting to find out what the effect is of applying the zero-inflation

upfront, thus before the main models. That way, an initial model could be applied to classify

a certain prediction in either the zero or non-zero class. For the non-zero predictions, the

main models could be applied. This may give different results compared to the post-model
zero-inflation tactic we applied.

64
Third, it would be equally interesting to find out the effect of calculating the daily loads as

the sums of the hourly loads. Although our hourly predictions are not yet on a suitable level,

in aggregate their accuracy might be better than for the current daily models.

Fourth, like Taieb & Hindman (2013) suggest, both daily and hourly predictions could

potentially be improved by creating a separate model for each hour. In their research,

significantly higher levels of accuracy were achieved for the hourly predictions than in the

present study. This somewhat contradicts the findings of Fay et al. (2003), as they proved

that sequential models generally outperform parallel models. However, the latter authors

also successfully applied a combination of sequential and parallel models to achieve their

optimal results.

Fifth, a similar extension, creating separate models for working days and non-working days
could improve results. This has been done by, among others, Neto & Fiorelli (2008).

Sixth, national holidays were not taken into account in this study. We expect such days to, in

terms of heating/cooling demand behave similar to weekend days, such that demands have

likely been overestimated when these holidays occurred on “working days”. A slight

increase in performance could result from correcting this.

Seventh, it would be interesting to apply ensemble techniques to this problem. Fan et al.

(2014) for instance applied a combination of (among others) multiple linear regression,

ARIMA, support vector regression, random forests and k-nearest neighbours to form an

ensemble model. This model significantly outperformed the separate models.

Eighth, taking into account the wind direction might add explanatory power, especially in

combination with the wind speed variable we included. Given that air infiltration is such a

large predictor of heating and cooling load – as Younes et al. (2012) showed – we believe that

including the wind direction, which is available in the KNMI data, might improve results.

Lastly, the study would of course greatly benefit from having more and especially more

complete data available. The large blocks that are missing in the set we used are, we believe,

an important cause of the hourly models underperforming. Also, some additional variables –

that were not monitored in the present dataset – could prove important. For instance, the

electricity consumption of the heat pumps was measured, but the electricity consumption of

65
the AHU’s was not. Thus, we could not subtract this from total power consumption to get as

close an approximation to occupancy as possible.

66
Literature list
Bauer, M., and Scartezzini, J.-L. (1998. ‘A simplified correlation method accounting for
heating and cooling loads in energy-efficient buildings’, Energy and Buildings, 27(2), 147–154.

Ben-Nakhi, A.E. and Mahmoud, M.A. (2004) ‘Cooling load prediction for buildings using

general regression neural networks’, Energy Conversion and Management 2004; 45:2127–41.

Bing, D. (2005) Baseline models of whole building energy consumption, ch. 3

Caudil, M. (1989) ‘Neural Network Primer Part I’, AI expert 02-1989

Cho, S.H., Kim, W.T., Tae, C.S. and Zaheeruddin, M. (2004) ‘Effect of length of measurement

period on accuracy of predicted annual heating energy consumption of buildings’, Energy

Conversion and Management, 45(18–19), 2867–28t78.

Datta, D., Tassou, S.A. and Marriott, D. (2000) ‘Application of neural networks for the

prediction of the energy consumption in a supermarket’, Proc. CLIMA 1–10.

Fan, C., Xiao, F. and Wang, S. (2014) ‘Development of prediction models for next-day

building energy consumption and peak power demand using data mining techniques’,

Applied Energy 127, 1–10.

Fan, C., Xiao, F. and Zhao, Y. (2017) ‘A short-term building cooling load prediction method

using deep learning algorithms’, Applied Energy 195, 222-233.

67
Fay, D., Ringwood, J.V., Condon, M. and Kelly, M. (2003) ‘24-h electrical load data sequential

or partitioned time series?’, Neurocomputing 55(3-4), 469–498.

Gelston, L. (2017) ‘Neural network training’. Accessed on July 7 2017 on

http://lauragelston.ghost.io/neural-network-training/.

González, P.A., and Zamarreño, J.M. (2005) ‘Prediction of hourly energy consumption in

buildings based on a feedback artificial neural network’, Energy and Buildings, 37(6), 595–601.

Kalogirou, S., Neocleous, C. and Schizas, C. (1997) ‘Building Heating Load Estimation Using

Artificial Neural Networks’, Proceedings of the 17th International Conference on Parallel

Architectures and Compilation Techniques 1–8.

Lai, F., Magoulès, F., and Lherminier, F. (2008) ‘Vapnik’s learning theory applied to energy

consumption forecasts in residential buildings’, International Journal of Computer Mathematics


85(0), 1563–1588.

McCaffrey, J.D. (2017) ‘Neural Network Momentum’. Accessed on July 17 2017 on


https://jamesmccaffrey.wordpress.com/2017/06/06/neural-network-momentum/.

Meyer, D. (2017) ‘Support Vector Machines: The Interface to libsvm in package e1071’.

Accessed on February 26 2017 on ftp://cran.r-

project.org/pub/R/web/packages/e1071/vignettes/svmdoc.pdf

68
Neto, A.H., and Fiorelli, F.A.S. (2008) ‘Comparison between detailed model simulation and

artificial neural network for forecasting building energy consumption’, Energy and Buildings

40(12), 2169–2176.

Newsham, G.R., and Birt, B.J. (2010) ‘Building-level Occupancy Data to Improve ARIMA-
based Electricity Use Forecasts’, BuildSys 2010, 13–18.

‘E-Handbook of Statistical Methods’ (2012) Accessed on July 5 2017 on

http://www.itl.nist.gov/div898/handbook/

Pérez-Lombard, L., Ortiz, J., and Pout, C. (2008.’ A review on buildings energy consumption

information’, Energy and Buildings 40(3), 394–398.

‘Neural Net’ (2017) Accessed on July 5 2017 on

https://docs.rapidminer.com/studio/operators/modeling/predictive/neural_nets/neural_net.h
tml.

Ramanathan, R., Engle, R. F., Granger, C.W.J., Vahid, F. and Brace, C. (1997) ‘Short-run
forecasts of electricity loads and peaks’, International Journal of Forecasting 13, 161–174.

Taieb, S.B. and Hyndman, R.J. (2013) ‘A gradient boosting approach to the Kaggle load

forecasting competition’.

Ürge-Vorsatz, D., Cabeza, L.F., Serrano, S., Barreneche, C. and Petrichenko, K. (2015)

‘Heating and cooling energy trends and drivers in buildings’, Renewable and Sustainable
Energy Reviews 2015; 41:85–98.

69
Wisse, K. & Abdalla, G. (2017), interviews.

Xue, X., Wang, S.W., Sun, Y.J., Xiao, F. (2014) ‘An interactive building power demand

management strategy for facilitating smart grid optimization’, Applied Energy 2014 ;116:297–

310.

Yang, Z. and Becerik-Gerber, B. (2017) ‘Assessing the impacts of real-time occupancy state

transitions on building heating/cooling loads’, Energy and Buildings 135, 201-211.

Yin, R., Xu, P., Piette, M.A. and Kiliccote, S. (2010) ‘Study on Auto-DR and pre-cooling of
commercial buildings with thermal mass in California’, Energy and Buildings, 42, 967–975.

Yuce, B., Li, H., Rezgui, Y., Petri, I., Jayan, B. and Yang, C. (2014) ‘Utilizing artificial neural

network to predict energy consumption and thermal comfort level: An indoor swimming

pool case study’, Energy and Buildings 80, 45–56.

Younes, C., Shdid, C.A., Bitsuamlak, G. (2012). Air infiltration through building envelopes: A

review. Journal of Building Physics. vol 35. pp 267. 2012

Zhao, H.X., and Magoulès, F. (2012) ‘A review on the prediction of building energy

consumption’, Renewable and Sustainable Energy Reviews 16(6), 3586–3592.

70
Appendix A: AHU, Toutdoor, Toutdoor setpoint data descriptives
Variable Unit of measurement

Date/time

Fan return steering signal %

Fan supply steering %

signal

Setpoint air pressure Pa

supply channel

Air pressure supply Pa

channel

Temperature air supply °C

RV return air %

Setpoint air blow-in °C

temperature

Valve cooling steering %

signal

Valve steam steering %

signal

Valve heating steering %

signal

Heat wheel steering %

signal heating

Air return temperature °C

Temperature cooling °C

verw-battery entrance

Temperature cooling °C

verw-battery exit

Table 16: AHU variables.

Variable Unit of measurement

Date/time

Outside temperature °C

Measured inside °C

temperature

Table 17: Toutdoor/Tindoor variables.

71
Variable Unit of measurement

Date/time

Outside temperature °C

Desired inside °C

temperature

Table 18: Toutdoor/Tindoor setpoint variables.

Most variables are self-explanatory, yet we will make some remarks regarding the heat

recovery system that is part of the AHU circuit.

First of all, the ‘heat wheel steering signal’ being a value > 0 indicates the heat recovery

system being active. Although we do not use the AHU data directly into our model, it does

lead to important insights: the heat wheel, when active, further deteriorates the potential

linear relationship between outside temperature and energy demand. As we saw earlier, a

certain amount of inertia is present by definition because of the building characteristics.

However, the heat recovery system complicates this relationship slightly more so. Heat being

recovered from the return flow does not ask any energy from the central heating circuit,

unlike the heating coil directly after the recovery system, thus introducing more nonlinearity

to the system. This notion also holds for the cooling circuit. The effect of this, however, is

relatively minor.

Once the value for the ‘Air pressure supply’ variable becomes greater than the ‘Setpoint air

pressure supply channel’ variable, this indicates the AHU’s being active. The same can be

deducted from the ‘Fan return steering signal’ being > 0. It is important to note however that

the fans of the AHU can be active, but no heating or cooling energy is being absorbed from

the central system. The latter can be deducted from the ‘Valve cooling steering signal’ and
‘Valve heating steering signal’ variables.

72
Appendix B: Neural network visualisation and description for daily
cooling model

Figure 21: Visualisation of neural network for the daily cooling model.

73
Hidden 1 TX: 0.689 TX-1: -0.356
======== Q: 0.211 TX-2: -0.216
UG: -0.193 Bias: 0.470
Node 1 (Sigmoid) Q-1: 0.185
---------------- Q-2: 0.156 Node 6 (Sigmoid)
Weekdag_vrijdag: -0.285 TG-1: 0.542 ----------------
Weekdag_zaterdag: 0.069 TG-2: 0.380 Weekdag_vrijdag: 0.221
Weekdag_zondag: 0.072 UG-1: 0.061 Weekdag_zaterdag: 0.041
Weekdag_dinsdag: -0.214 UG-2: 0.060 Weekdag_zondag: 0.135
Weekdag_woensdag: -0.191 FG-1: -0.098 Weekdag_dinsdag: 0.174
Weekdag_donderdag: -0.253 FG-2: -0.072 Weekdag_woensdag: 0.200
Sum of Residual_elec: - Sum of Residual_elec-1: - Weekdag_donderdag: 0.171
0.421 0.304 Sum of Residual_elec:
FG: 0.026 Sum of Residual_elec-2: - 0.220
TG: -0.333 0.402 FG: 0.083
TX: -0.330 TX-1: 0.532 TG: 0.140
Q: -0.146 TX-2: 0.373 TX: 0.223
UG: 0.131 Bias: -0.742 Q: 0.165
Q-1: -0.072 UG: -0.036
Q-2: -0.005 Node 4 (Sigmoid) Q-1: 0.079
TG-1: -0.185 ---------------- Q-2: 0.097
TG-2: -0.093 Weekdag_vrijdag: 0.140 TG-1: 0.114
UG-1: 0.036 Weekdag_zaterdag: 0.166 TG-2: 0.061
UG-2: 0.018 Weekdag_zondag: 0.173 UG-1: 0.009
FG-1: 0.050 Weekdag_dinsdag: 0.144 UG-2: -0.042
FG-2: 0.085 Weekdag_woensdag: 0.180 FG-1: -0.004
Sum of Residual_elec-1: Weekdag_donderdag: 0.131 FG-2: 0.030
0.008 Sum of Residual_elec: - Sum of Residual_elec-1: -
Sum of Residual_elec-2: 0.035 0.057
0.135 FG: 0.086 Sum of Residual_elec-2: -
TX-1: -0.258 TG: -0.035 0.050
TX-2: -0.142 TX: -0.094 TX-1: 0.118
Bias: 0.280 Q: 0.056 TX-2: 0.107
UG: -0.037 Bias: -0.308
Node 2 (Sigmoid) Q-1: 0.049
---------------- Q-2: 0.112 Node 7 (Sigmoid)
Weekdag_vrijdag: 0.090 TG-1: 0.006 ----------------
Weekdag_zaterdag: 0.059 TG-2: -0.032 Weekdag_vrijdag: 0.170
Weekdag_zondag: 0.140 UG-1: -0.037 Weekdag_zaterdag: 0.150
Weekdag_dinsdag: 0.152 UG-2: 0.032 Weekdag_zondag: 0.148
Weekdag_woensdag: 0.157 FG-1: 0.117 Weekdag_dinsdag: 0.165
Weekdag_donderdag: 0.084 FG-2: 0.117 Weekdag_woensdag: 0.157
Sum of Residual_elec: - Sum of Residual_elec-1: - Weekdag_donderdag: 0.170
0.003 0.076 Sum of Residual_elec: -
FG: 0.094 Sum of Residual_elec-2: - 0.020
TG: -0.131 0.021 FG: 0.131
TX: -0.124 TX-1: 0.013 TG: -0.078
Q: 0.130 TX-2: 0.024 TX: -0.001
UG: 0.002 Bias: -0.241 Q: 0.104
Q-1: 0.134 UG: -0.051
Q-2: 0.119 Node 5 (Sigmoid) Q-1: 0.087
TG-1: 0.027 ---------------- Q-2: 0.095
TG-2: 0.031 Weekdag_vrijdag: -0.421 TG-1: 0.014
UG-1: -0.031 Weekdag_zaterdag: 0.048 TG-2: -0.016
UG-2: 0.033 Weekdag_zondag: 0.100 UG-1: -0.004
FG-1: 0.178 Weekdag_dinsdag: -0.251 UG-2: -0.025
FG-2: 0.161 Weekdag_woensdag: -0.343 FG-1: 0.099
Sum of Residual_elec-1: - Weekdag_donderdag: -0.372 FG-2: 0.167
0.114 Sum of Residual_elec: - Sum of Residual_elec-1: -
Sum of Residual_elec-2: - 0.612 0.095
0.032 FG: 0.005 Sum of Residual_elec-2: -
TX-1: -0.024 TG: -0.511 0.024
TX-2: 0.075 TX: -0.490 TX-1: 0.052
Bias: -0.159 Q: -0.228 TX-2: 0.072
UG: 0.154 Bias: -0.175
Node 3 (Sigmoid) Q-1: -0.158
---------------- Q-2: -0.076 Node 8 (Sigmoid)
Weekdag_vrijdag: 0.444 TG-1: -0.329 ----------------
Weekdag_zaterdag: -0.045 TG-2: -0.248 Weekdag_vrijdag: 0.160
Weekdag_zondag: -0.001 UG-1: -0.029 Weekdag_zaterdag: 0.143
Weekdag_dinsdag: 0.252 UG-2: 0.031 Weekdag_zondag: 0.173
Weekdag_woensdag: 0.560 FG-1: 0.073 Weekdag_dinsdag: 0.177
Weekdag_donderdag: 0.387 FG-2: 0.024 Weekdag_woensdag: 0.126
Sum of Residual_elec: Sum of Residual_elec-1: Weekdag_donderdag: 0.145
0.738 0.126 Sum of Residual_elec: -
FG: 0.003 Sum of Residual_elec-2: 0.070
TG: 0.622 0.229 FG: 0.121

74
TG: -0.009 Sum of Residual_elec-2: Sum of Residual_elec:
TX: -0.053 0.008 0.088
Q: 0.132 TX-1: 0.065 FG: 0.100
UG: -0.032 TX-2: -0.004 TG: 0.048
Q-1: 0.168 Bias: -0.257 TX: 0.023
Q-2: 0.113 Q: 0.096
TG-1: 0.030 Node 11 (Sigmoid) UG: -0.005
TG-2: 0.044 ----------------- Q-1: 0.067
UG-1: 0.011 Weekdag_vrijdag: 0.211 Q-2: 0.062
UG-2: -0.024 Weekdag_zaterdag: 0.102 TG-1: 0.085
FG-1: 0.181 Weekdag_zondag: 0.166 TG-2: 0.047
FG-2: 0.154 Weekdag_dinsdag: 0.186 UG-1: -0.000
Sum of Residual_elec-1: - Weekdag_woensdag: 0.119 UG-2: 0.006
0.100 Weekdag_donderdag: 0.166 FG-1: 0.075
Sum of Residual_elec-2: Sum of Residual_elec: - FG-2: 0.072
0.002 0.047 Sum of Residual_elec-1: -
TX-1: -0.004 FG: 0.079 0.025
TX-2: 0.040 TG: -0.018 Sum of Residual_elec-2: -
Bias: -0.187 TX: 0.020 0.007
Q: 0.112 TX-1: 0.110
Node 9 (Sigmoid) UG: -0.021 TX-2: 0.075
---------------- Q-1: 0.141 Bias: -0.252
Weekdag_vrijdag: 0.158 Q-2: 0.080
Weekdag_zaterdag: 0.120 TG-1: -0.037 Node 14 (Sigmoid)
Weekdag_zondag: 0.197 TG-2: 0.042 -----------------
Weekdag_dinsdag: 0.168 UG-1: -0.023 Weekdag_vrijdag: 0.066
Weekdag_woensdag: 0.134 UG-2: 0.012 Weekdag_zaterdag: 0.048
Weekdag_donderdag: 0.148 FG-1: 0.071 Weekdag_zondag: 0.156
Sum of Residual_elec: - FG-2: 0.163 Weekdag_dinsdag: 0.102
0.003 Sum of Residual_elec-1: - Weekdag_woensdag: 0.177
FG: 0.124 0.002 Weekdag_donderdag: 0.055
TG: -0.052 Sum of Residual_elec-2: - Sum of Residual_elec: -
TX: -0.023 0.048 0.008
Q: 0.104 TX-1: -0.006 FG: 0.110
UG: -0.043 TX-2: 0.021 TG: -0.133
Q-1: 0.098 Bias: -0.202 TX: -0.103
Q-2: 0.163 Q: 0.083
TG-1: -0.038 Node 12 (Sigmoid) UG: 0.036
TG-2: 0.026 ----------------- Q-1: 0.080
UG-1: -0.007 Weekdag_vrijdag: 0.016 Q-2: 0.139
UG-2: -0.037 Weekdag_zaterdag: 0.099 TG-1: -0.066
FG-1: 0.168 Weekdag_zondag: 0.099 TG-2: 0.009
FG-2: 0.172 Weekdag_dinsdag: 0.033 UG-1: -0.028
Sum of Residual_elec-1: - Weekdag_woensdag: 0.083 UG-2: -0.023
0.042 Weekdag_donderdag: -0.002 FG-1: 0.175
Sum of Residual_elec-2: - Sum of Residual_elec: - FG-2: 0.119
0.071 0.089 Sum of Residual_elec-1: -
TX-1: 0.012 FG: 0.098 0.046
TX-2: 0.020 TG: -0.182 Sum of Residual_elec-2: -
Bias: -0.156 TX: -0.112 0.005
Q: 0.018 TX-1: -0.037
Node 10 (Sigmoid) UG: 0.045 TX-2: 0.112
----------------- Q-1: 0.035 Bias: -0.148
Weekdag_vrijdag: 0.190 Q-2: 0.147
Weekdag_zaterdag: 0.100 TG-1: -0.079
Weekdag_zondag: 0.152 TG-2: 0.059 Output
Weekdag_dinsdag: 0.182 UG-1: -0.049 ======
Weekdag_woensdag: 0.163 UG-2: -0.016
Weekdag_donderdag: 0.126 FG-1: 0.151 Regression (Linear)
Sum of Residual_elec: FG-2: 0.156 -------------------
0.006 Sum of Residual_elec-1: - Node 1: 0.290
FG: 0.052 0.046 Node 2: 0.291
TG: 0.010 Sum of Residual_elec-2: - Node 3: -1.237
TX: -0.023 0.034 Node 4: 0.158
Q: 0.118 TX-1: -0.015 Node 5: 0.672
UG: -0.026 TX-2: -0.015 Node 6: -0.092
Q-1: 0.052 Bias: -0.075 Node 7: 0.195
Q-2: 0.110 Node 8: 0.259
TG-1: -0.036 Node 13 (Sigmoid) Node 9: 0.233
TG-2: -0.011 ----------------- Node 10: 0.150
UG-1: -0.002 Weekdag_vrijdag: 0.208 Node 11: 0.180
UG-2: 0.023 Weekdag_zaterdag: 0.077 Node 12: 0.167
FG-1: 0.142 Weekdag_zondag: 0.143 Node 13: 0.082
FG-2: 0.127 Weekdag_dinsdag: 0.171 Node 14: 0.246
Sum of Residual_elec-1: - Weekdag_woensdag: 0.147 Threshold: -0
0.007 Weekdag_donderdag: 0.138

75

S-ar putea să vă placă și