You are on page 1of 3

Download PDF Search ScienceDirect

Journals & Books Register Sign in

Outline Recommended articles


Energy and Buildings
Highlights Energetic study of a Trombe wall system under…
Volume 80, September 2014, Pages 218-230
Abstract Energy and Buildings, Volume 80, 2014, pp. 302-308

Purchase PDF View details


Keywords

Nomenclature Effect of dynamic characteristics of building Assessment of the dynamic thermal performan…
1. Introduction Energy and Buildings, Volume 72, 2014, pp. 361-370
envelope on thermal-energy performance in Purchase PDF View details
2. Purpose of the research

3. The buildings
winter conditions: In field experiment Using suite energy-use and interior condition …
4. The monitoring setup Anna Laura Pisello , Franco Cotana, Andrea Nicolini, Cinzia Buratti Energy and Buildings, Volume 80, 2014, pp. 184-194

5. Results and discussion Purchase PDF View details


Show more
6. Conclusions 1 2 Next
Share Cite
Acknowledgments

References https://doi.org/10.1016/j.enbuild.2014.05.017 Get rights and content Citing articles (46)


Show full outline

Article Metrics
Figures (12)
Highlights Citations
• Two buildings with same stationary properties and different dynamics
Citation Indexes: 46
are built.
Captures
• Results of thermal-energy indoor–outdoor continuous monitoring
campaign are analyzed. Exports-Saves: 3
Readers: 119
• Comparative dynamic behavior of the two buildings is evaluated in
winter conditions.
View details
Show all figures
• Energy behavior is equivalent. Field conductance, humidity and
temperature differ.
Tables (3)
• Impact of building dynamic properties should be considered even in
Table 1
Table 2 winter condition.
Table 3

Abstract
Research about building strategies for energy saving is taking a growing interest
worldwide. New solutions for building envelopes require increasingly
sophisticated investigation to analyze the thermal-energy in-field dynamic
response of constructions. Moreover, in-lab experiments are difficultly able to
represent real environment boundary conditions. In this paper, effect of dynamic
properties of building opaque envelope is investigated in winter conditions
through an extensive continuous monitoring campaign of a dedicated
experimental field. The field consists of: two full-scale buildings, an outdoor
weather station, and two indoor microclimate stations. The two buildings were
designed with the same stationary envelope characteristics, but different envelope
technologies, materials and, therefore, different dynamic properties. Nevertheless,
following Italian regulations about building energy performance in winter, they
should behave the same. With the purpose to verify this hypothesis, a continuous
long-term comparative monitoring is developed. The results only partly confirm
that simplified hypothesis. In fact, in winter, while the two buildings exhibit
equivalent long-term energy behavior, a weak difference is registered in terms of
air temperature in free-floating, transient, and operative HVAC systems’ regime.
Additionally, non-negligible discrepancies are registered in terms of mean radiant
temperature, indoor humidity and internal–external envelope surface
temperature, given the different transpiration rate of the two envelope systems
and different solar reflectance values of the roofs.

Previous article in issue Next article in issue

Keywords
Building continuous monitoring; Energy efficiency in buildings; Indoor thermal
behavior; Building envelope; Building thermal-energy performance; In-field
envelope conductance and transmittance measurement; Dynamic methods for
building thermal-energy assessment

Nomenclature
r
Pearson correlation coefficient, –

RH
relative humidity, %

T
temperature, K

Tair
indoor air temperature, °C

Tmr
indoor mean radiant temperature, °C

Top
indoor operative temperature, °C

TR
Test-Room, monitored prototype building

TR-1 and TR-2


Test-Room 1 and 2, respectively

VC
coefficient of variation, –

σ
standard deviation

1. Introduction
The research interest focused on buildings’ energy efficiency, passive solar
architecture, innovative optimization strategies and retrofit interventions has
received a growing interest in last decades [1], given the acknowledged key role of
constructions to reduce global environment impact in terms of fossil fuels’
depletion and carbon emissions [2]. In particular, many research efforts focused on
elaborating integrated multicriteria tools aimed at optimizing the overall
buildings’ energy performance through: (i) envelopes’ improvement, (ii) reduction
of heating and cooling requirement, and (iii) innovative control strategies [3]. In
order to improve the overall thermal-energy performance of opaque and
transparent components of buildings’ envelope, several scientific contributions
focused on indoor thermal-energy monitoring of specific case studies with the
purpose to quantify benefits and possible penalties of tested retrofit solutions
when exposed to different climate conditions [4]. For instance, a wide investigation
concerned the implementation of innovative envelope materials and coatings on
case study buildings, after an in-lab analysis and optimization of the tested
solutions [5]. Particular attention was paid to the effect of cool roofs and cool
coatings to reduce building energy requirement for cooling in different
climatological contexts, through experimental monitoring and calibrated-
validated dynamic simulation [6], [7]. In particular, Kolokotsa et al. in [7] elaborated
a numerical and experimental analysis of a cool roof application on a research
office building in Crete, Greece. The analysis was carried out through continuous
indoor–outdoor environmental monitoring and validated dynamic simulation, in
order to extend the experimental results to a number of variations in boundary
conditions and for longer periods than the experiment duration. The need for
dynamic simulation modeling was also to analyze the influence of several
innovative passive techniques [8] impacting the indoor thermal performance of the
monitored areas, e.g. occupants-based strategies [9].

In order to develop rigorous thermal-energy continuous monitoring, important


researches during last decades also focused on the performance analysis of
dedicated full-scale test-buildings, which allowed higher dynamic parameters’
control than existing occupied buildings. In particular, a key research contribution
was produced by the development of PASSYS research project, where several
important papers were published in the field of experimental analysis of thermal
properties through outdoor test cell facilities [10] since 1993. Then, the necessity
to discuss further around dynamic thermal behavior of buildings, and building
components in particular, was focused by the PASLINK network [11]. They
provided significant improvements in the field of both testing and analysis
techniques, also guiding other researchers in creating new test facilities, such as
the research field presented in this work [12]. In particular, Strachan and Vandeele
in [12] presented a series of case studies concerning specific building components,
in order to analyze and to better model their dynamic field behavior through
sophisticated computer modeling techniques. Additionally, they showed that
laboratory conditions could produce lower uncertainties, but the only outdoor
continuous monitoring facilities allow the investigation of real wide operational
ranges of dynamic conditions’ variability. In this view, Baker and van DijK in [13]
reviewed both stationary and dynamic methodology for assessing building thermal
performance. In particular, they applied parameter identification in order to
investigate dynamic effects imputable to heat accumulation. The transient
mathematical model led to the development of new softwares for parameter
identifications, which implementation could be useful in order to compare the
research of this paper with respect to other key contributions by PASLINK
network [12], [13], as future development of this work. Additionally, Jiménez and
Madsen, starting from the results of outdoor testing, presented an overview of
models aimed at predicting thermal characteristics of buildings and building
components [14]. In particular, they outlined the main differences between linear,
non-linear, and time-invariant approaches. Therefore, they applied some of the
presented models to the case study of the PASLINK experimental field, with the
purpose to estimate the thermal transmission coefficient and the solar
transmittance of the tested wall (the periodically modifiable part of the opaque
envelope) in the cell. This series of contributions shows how such outdoor test
facility could also be useful for identifying, and therefore optimizing, the
performance of each building component, such as the experimental wall of [10],
[11], [12], [13], [14], [15], that was also used to present the procedure for using
IDENT in MATLAB [15]. Additionally, the necessity of an up-to-date research
network sharing the same purpose demonstrated the scientific need of this kind
of sophisticated multivariable experiments [16]. In this view, the purpose of these
full scale buildings (i.e. the Test-Room experiment presented in this work) mainly
consists of the thermal-energy and humidity analysis of the behavior of real
buildings, i.e. with common envelope multilayer structure and HVAC systems. In
this view, this paper in particular concerns this experimental investigation with
varying winter conditions, showing the importance of dynamically variable
boundaries in determining building indoor environmental quality.

Starting from these key research contributions [10], [11], [12], [13], [14], [15], [16],
other interesting applications of small-scale buildings were carried out in order to
simulate and investigate the role of buildings’ envelope on the thermal behavior of
the urban canopy [17] and local climate phenomena such as urban heat island
effect and its huge impact on buildings’ energy performance [18]. Thanks to the
implementation of such a sophisticated experimental setup, important
enhancements were carried out to investigate building envelope properties and to
optimize innovative materials for energy saving. In particular, the role of reflective
coatings for both interior and exterior cladding of building envelope was
investigated by Joudi et al. in [19]. In this study, three monitored steel clad test-
cabins were used to investigate the role of infrared-reflective coatings under
various conditions, which results were also used to evaluate the year-round energy
use of the same cabins. A detailed analysis concerning several innovative roof
technologies was carried out by D’Orazio et al. in [4], where the experimental setup
consisted of a test building equipped with six roof modules. Each module
represented a different technology, i.e. green roof, traditional clay tile roof, etc. and
the overall setup was monitored in terms of roof thermal behavior and outdoor
weather conditions.

Notable enhancements in the field of house-sized dedicated building monitoring


were carried out in order to study the dynamic thermal performance of buildings’
envelopes such as the thermal inertia characteristic. In particular, the investigation
of PCMs’ (Phase Change Materials) effect produced important results thanks to the
continuous monitoring of small house-sized cubicles carried out in [20]. Castellón
et al. in [20] presented the experimental setup consisting of nine small house-size
cubicles made with brick or concrete, where PCM applications were tested
through continuous monitoring of free-floating conditions and operative HVAC
system conditions. The monitoring equipment consisted of indoor and surface
temperature sensors in each test-cell, while the weather data were available from
an autonomous station located nearby. The internal dimensions of the prototypes
(2.4 m × 2.4 m × 2.4 m) were established in order to be big enough to be
representative, but small enough to be economically feasible, and the mutual
position of the test-cells was established in order to avoid inter-building effects
affecting the measurements [21]. The same PCM applications were studied by
Bontemps et al. in [22] which purpose was to evaluate the role of PCM inclusion
into a wall in terms of indoor temperature and thermal flux through the same
wall. To this aim an outdoor test cell was constructed and it consisted of two
juxtaposed rooms separated by glass bricks with or without PCMs’ inclusion. The
experimental setup was monitored through flux-meters and thermocouples which
data were used to calibrate and validate one-dimensional numerical model with
the final purpose to characterize several PCMs’ implementations.

The growing interest in evaluating new high performance envelope technologies


and materials guided through the elaboration of the experimental campaign
presented in this paper. The concept of the two full-scale test-buildings, designed
in order to represent typical Italian construction practice in terms of materials,
geometry and technical properties of the envelope and of the HVAC technologies
[23], was carried out by following the recent regulation about energy efficiency in
buildings. The Test-Rooms were designed and built in 2012. Each Test-Room was
equipped with a complete indoor monitoring station connected to a dedicated
outdoor weather station. Differently with respect to the previous experimental
campaigns described before, these buildings were constructed with traditional
materials and complete envelope solutions, such as real doors and windows,
structural systems, and multilayer walls-roof-ceiling. The coupled parallel analysis
concerning indoor–outdoor environment of both the Test-Rooms was mainly
conceived to investigate the comparative hygrothermal and energy behavior of
such cabins in winter conditions. Therefore, the role of important dynamic
features, usually neglected in Italian regulation for building design in winter
conditions, e.g. the active and passive contribution of envelope reflectance [24],
[25], [26], [27], at single building and larger scale position of the insulation panel,
transpiring characteristic of materials, is analyzed [10], [11], [12], [13], [14], [15].

2. Purpose of the research


The purpose of this experimental research is the evaluation of different envelope
typologies, presenting the same stationary thermal properties, through continuous
monitoring analysis of combined indoor and outdoor environmental conditions.
In particular, the specific characterization of each Test-Room is designed in order
to represent: firstly, a traditional Italian construction technique, secondly, a more
innovative construction technique, commonly used also as building energy
retrofit. The first one (Test-Room 1), as specifically described in the following
section of this work, consists of a double brick layer façade, with the insulation
panel positioned in the wall cavity and a low-sloped roof structure covered by
natural red clay tiles. The latter Test-Room (Test-Room 2) presents an external
insulated façade and a horizontal roof with external insulation panel and bitumen
waterproof membrane. Given that the structural system of both the buildings
consists of reinforced concrete structure, the façade systems were designed to be
periodically modified for research purpose, in order to evaluate and optimize
several configurations. All the tested solutions well represent realistic
configurations of buildings for the choice of materials and technologies, structural
system, and positioning of the Test-Rooms. The location is selected in order to
avoid differential shading elements like trees or other buildings located in close
proximity, responsible for differential thermal-energy behavior of the Test-Rooms
and other inter-building effects. Additionally, the main architecture parameters
are designed to reproduce realistic conditions. In particular, the Window Wall
Ratio (WWR) corresponds to 17% for the South façade and the ratio between the
envelope surface and the volume included within the envelope surface
corresponds to about 1.7 for both the Test-Rooms, which could reproduce Italian
single-family residential buildings [28], [29]. The ratio between the window area
and the flooring surface is 19%, according to Italian regulation, i.e. higher than
12.5%.

In this work the results of this permanent experiment concerning the data
analysis of winter continuous thermal-energy monitoring campaign are presented.
The paper is organized as follows: Section 3 presents the building physics of the
Test-Rooms and Section 4 concerns the detailed description of the continuous
monitoring setup. In Section 5 the results of the winter continuous thermal-
energy monitoring are reported in order to characterize the parallel behavior of
the two Test-Rooms during both free-floating regime, transient regime and
operative HVAC system regime in winter conditions [30]. The main conclusions
are drawn in the final section of this work.

3. The buildings
The experimental platform, compared to real building measurement campaign,
allows continuous and permanent monitoring of hygrothermal-energy
performance of the Test-Rooms and of the weather conditions with high
controlled boundary conditions. In this view, the experimental field allows to
investigate the role of several envelope techniques and technologies in the
thermal-energy performance of the building prototypes without any affection by
random parameters, which usually affect the reliability of the continuous
monitoring of full-scale buildings with on-going operations [6], [17]. Nevertheless,
the simplification of the geometry and of the boundary conditions implies
obvious discrepancy between the Test-Room behavior and real building's behavior.
In this view, the analysis presented in this paper concerns the comparative
assessment of the thermal-energy performance of these two geometrically and
functionally equivalent buildings.

3.1. Construction and materials


The two case study buildings consist of two prototypes, which characterization is
reported in Fig. 1, located in front of the Engineering School of University of
Perugia, Italy. The positioning was studied in order to avoid mutual shading
during the day, in compliance with the land availability of the campus. The
southern Test-Room, Test-Room 1 (Fig. 1) just shades the Northern one, Test-
Room 2, at the very beginning of the day (before 10:00 a.m.) in winter months. A
rectangular double shutter window with wood frames was installed on the South
façade of each Test-Room; the window was in-lab characterized through hot-box
methodology [31]. The same typology of rectangular armored door was installed in
the North façade of each Test-Room.

Download : Download full-size image

Fig. 1. Northern and Southern views of the Test-Rooms and their architecture.

Particular attention was paid to design the opaque envelope systems where, as
previously mentioned, the objective was to represent innovative and traditional
Italian construction typologies characterized by the same thermal stationary
properties for walls, roof and ceiling. Technical features of the materials used for
all the envelope components of each Test-Room are reported in Table 1.

Table 1. Characterization of the envelope components of the Test-Rooms.

Test-Room 1_External wall

1. Brickwork, outer Thickness: Conductivity: Properties [32]


leaf 0.12 m 0.41 W/m K Thermal transmittance (surface-to-
surface): 0.29 W/m2 K
2. Plasterboard 0.01 m 0.16 W/m K
Decrement factor: 0.465 [–]
3. EPS insulation 0.09 m 0.04 W/m K Time shift: 6.78 h

4. Brickwork, inner 0.25 m 0.31 W/m K Periodic thermal transmittance:

leaf 0.164 W/m2 K


Aeral heat capacity (inner side):
5. Gypsum 0.02 m 0.40 W/m K
22.1 kJ/m2 K
plastering
Aeral heat capacity (outer side):
82.1 kJ/m2 K

Test-Room 2_External wall

1. Plaster dense Thickness: Conductivity: Properties [32]


0.02 m 0.50 W/m K Thermal transmittance (surface-to-
surface): 0.29 W/m2 K
2. EPS insulation 0.09 m 0.04 W/m K
Decrement factor: 0.973 [–]
3. Brickwork, inner 0.30 m 0.27 W/m K Time shift: 1.46 h
leaf Periodic thermal transmittance:

4. Gypsum 0.02 m 0.40 W/m K 0.388 W/m2 K

plastering Aeral heat capacity (inner side):


24.92 kJ/m2 K
Aeral heat capacity (outer side):
62.35 kJ/m2 K

Test-Room 1_Roof

1. Clay tile Thickness: Conductivity: Properties [32]


0.015 m 1.00 W/m K Thermal transmittance (surface-to-
surface): 0.25 W/m2 K
2. Mineral wool 0.015 m 0.04 W/m K
Decrement factor: 0.311 [–]
insulation
Time shift: 8.78 h
3. Air gap 0.05 m Thermal resistance: Periodic thermal transmittance:
(ventilation layer) 0.23 m2 K/W 0.088 W/m2 K

4. Mineral wool 0.08 m Conductivity: Aeral heat capacity (inner side):

insulation 0.038 W/m K 19.90 kJ/m2 K


Aeral heat capacity (outer side):
5. Aerated concrete 0.20 m 0.16 W/m K
27.40 kJ/m2 K
slab

6. Gypsum 0.015 m 0.40 W/m K


plastering

Test-Room 2_Roof

1. Bitumen sheet Thickness: Conductivity: Properties [32]


0.01 m 0.23 W/m K Thermal transmittance (surface-to-
surface): 0.25 W/m2 K
2. Mineral wool 0.10 m 0.04 W/m K
Decrement factor: 0.296 [–]
insulation
Time shift: 8.88 h
3. Aerated concrete 0.20 m 0.16 W/m K Periodic thermal transmittance:
slab 0.075 W/m2 K

4. Gypsum 0.015 m 0.40 W/m K Aeral heat capacity (inner side):

plastering 35.30 kJ/m2 K


Aeral heat capacity (outer side):
26.90 kJ/m2 K

Test-Room 1 and Test-Room 2_Ground floor

1. Linoleum Thickness: Conductivity: Properties [32]


0.004 m 0.17 W/m K Thermal transmittance: 0.30 W/m2 K
Decrement factor: 0.138 [–]
2. Glass fiber slab 0.10 m 0.04 W/m K
Time shift: 10.78 h
3. Cast concrete 0.30 m 1.13 W/m K Periodic thermal transmittance:
0.046 W/m2 K
Aeral heat capacity (inner side):
9.30 kJ/m2 K
Aeral heat capacity (outer side):
125.20 kJ/m2 K

Glazing systems: optical and thermal features of the

Visible light Solar radiation

Transmittance [%] 66 Transmittance [%] 39

Outside reflectance 26 Reflectance [%] 41


[%]

Inside reflectance 23 Absorbance [%] 20


[%]

Color rendering 96 –
index

Solar heat gain coefficient g [%]: 42

Thermal transmittance U [W/m2 K]: 1.3

Particular attention was paid in order to minimize the thermal bridges’ effect. To
this aim, the insulation panels were installed by maintaining the insulation
continuity along the walls and ceilings. Therefore, a rigid mineral wall layer was
positioned under the first layer of blocks. The same was operated for the windows,
where a wood slat was positioned under the window frame, in order to interrupt
the continuity of the window sill, which usually represents a relevant thermal
bridge. The two windows have rectangular shapes measuring 1.23 m × 1.48 m. The
pine wood frame presents 69 mm × 80 mm thick section, while the two shutters
have 69 mm × 78 mm thick section. The shutters are kept closed for the entire
duration of this continuous monitoring, in order to focus the analysis on the
opaque envelope characteristics. Between the frame and the shutters, two series of
rubber gaskets are installed, in order to minimize the infiltrations. The glass
panels consist of high performance double glazing systems where the external
panel (4 mm thick) presents a low-emissivity treatment on its internal side. The
camera, which is 15 mm thick, is air filled; the internal glass panel is a float glass
4 mm thick. The thermal features of the double glazing are reported in Table 1.
The solar protection device is represented by aluminum external venetian
shutters, which are mostly used in Italy for both solar shading and security
purpose [33].

3.2. The energy systems


The two monitored Test-Rooms present identical energy plants, both in terms of
lighting system and heating–cooling equipment. The lighting system consists of a
simple fluorescent lamp, roof surface mounted, which lighting energy
corresponds to 3.3 W/m2 (100 lux). The radiant fraction corresponds to 0.72, while
the visible fraction is 0.18. It is turned off for the entire duration of the study,
given that the periodical inspections are usually carried out when daylight is
available.

The air-conditioning system consists of a heat pump with inverter, which external
units are positioned over the roof of each Test-Room, and the internal units
consist of a split for each room. The nominal capacity of the heat pumps is 2.0 kW
and 2.5 kW for cooling and heating, respectively. The Energy Efficiency ratio
corresponds to 3.63 and the Coefficient of Performance corresponds to 4.24.

4. The monitoring setup


The coupled indoor–outdoor continuous monitoring was performed through a set
of sensors which were continuously operating since November 2012. The overall
scheme of the equipment consisted of (i) sensors, (ii) data-loggers, (iii) converters,
and (iv) indoor downloading station. The monitoring setup comprehends a series
of sensors for both microclimate and meteorological continuous monitoring,
together with energy meters which are connected to data-loggers in order to
measure and memorize the monitored data for a maximum duration of 67 days
(with electricity) and 18 days (with autonomous batteries). Each data-logger is
connected to a converter in order to transfer the signal from RS232 line to RS485
line, commonly used in balanced digital multipoint systems, which is able to
connect different equipment up to 1200 m far apart. The final re-conversion of the
signal happens at the indoor downloading station, positioned in the university
department, through another final converter. All the measurements are carried out
every 20 s by all the sensors, except for the wind sensors that measure every 5 s.
The average, maximum, minimum and the standard deviation values are
calculated every 10 min by the same data-logging system.

Given the intrinsic flexibility of the experimental setup and the possibility to
integrate even further applications and technologies, additional temporary
monitoring and analysis stations are periodically installed within the Test-Room
experimental field. For example, equivalent independent monitoring and
datalogging stations are installed for lighting and acoustic analysis [28], [29], or
thermal comfort assessment [30] within each room.

4.1. Indoor–oudoor continuous monitoring setup


The Test-Rooms present two tween indoor stations (Fig. 2a) which functionality is
aimed at monitoring (i) the indoor air thermal conditions, (ii) the envelope
performance in terms of superficial temperature of walls and roof, (iii) the
reflected radiation by the roof for cool roof applications [5], and (iv) the energy
consumption of the heat pump [25]. Table 2 reports the characterization of each
sensor and the measured parameters.

Download : Download full-size image

Fig. 2. Indoor–Outdoor microclimate monitoring equipment.

Table 2. Characterization of the indoor–outdoor monitoring stations and positioning.

Monitoring indoor station installed in each Test-Room

Sensors Monitored parameters Measurement features

Hot wire anemometer, turbolence intensity probe. Air velocity [m/s] Accuracy (10 ÷ 30 °C):
Position 1a 0 ÷ 0.5 m/s: ±5 cm;
0.5 ÷ 1.5 m/s: ±10 cm;
>1.5 m/s: 4%

Turbolence intensity [%] Sensitivity: 0.1 m/s

Global thermometer: probe for radiant temperature Mean radiant Accuracy: 0.15 °C (to
measurement. Position 1a temperature [°C] 0 °C)

Thermohygrometric probe. Position 1a Air temperature [°C] Accuracy: 0.15 °C (to


0 °C)

Air relative humidity [%] Accuracy: 3%


(T = 15 ÷ 45 °C)

N.4 probes for temperature measurement with Surface temperature of Accuracy: 0.15 °C (to
contact on the North side wall, where 2 probes are internal and external 0 °C)
installed in the internal side and 2 probes are side of the façade for
installed in the external side of the façade. Position transmittance
2b assessment [°C]

N.2 probes for temperature measurement with Surface temperature of Accuracy: 0.15 °C (to
contact on the roof, where 1 probe is installed in the internal and external 0 °C)
internal side and 1 probe is installed in the external side of the roof [°C]
side of the roof. Position 3c

N.2 thermal flux probe, where 1 probe is installed in Thermal flux through Sensitivity: 50 μV/m2
the internal side and 1 probe is installed in the the north external wall
Resolution: 0.1 W/m2
external side of the roof. Position 4d and through the roof
[W/m2]

Thermopile global radiation sensor (piranometer Global radiation Spectral range:


downward oriented). Position 5e reflected by the roof 305 ÷ 2800 nm
[W/m2]
Uncertainty: 10% daily

Second class WMO


(ISO 9060)

Resolution: 0.1 W/m2

Energy meter. Position 6f Energy consumption Minimum working


[kWh] current: 40 mA
Measuring voltage:
230 V ± 10%
Frequency: 50/60 Hz
Pulse output SO value:
1000 pulses/kWh

Operating
temperature:
−20 ÷ +55 °C

Outdoor Weather stationg

Sensors Monitored parameters Measurement features

Air speed sensor (cup) Wind velocity [m/s] Accuracy: 0.1 m/s
sensitivity: 0.4 m/s

Wind direction (vane) Prevailing wind Accuracy: 1%


direction, wind direction Sensitivity: 0.4 m/s
[°]

Thermohygrometer sensor Dry bulb temperature, Accuracy: 0.1 °C (0 °C)


Tout [°C]

Air relative umidity [%] Accuracy: 1.5%


(5 ÷ 95%, 23 °C)

Direct radiometer, with sunshine duration sensor Sunshine duration


(referred to a certain
threshold) [0–1]

Direct radiation from the Accuracy: 5% + 5 W/m3


sun [W/m2]

Thermopile global radiation sensor (piranometer Global solar irradiance Spectral range:
upward oriented) [W/m2] 305 ÷ 2800 nm

Uncertainty: 10% daily

Second class WMO


(ISO 9060)

Electric rain gauge Rain fall [mm] Resolution:


0.2 mm/imp

Accuracy:
0 ÷ 1 mm/min: 1%;
1 ÷ 3 mm/min: 2%;
3 ÷ 5 mm/min: 4%
5 ÷ 10 mm/min: 8%

a
Position 1. Center of each room, h = 1.10 m, distance from each wall, d = 1.5 m.

b
Position 2. Contact temperature probes positioned (i) on North facing wall, on both internal and
external side, h = 1.4 m; (ii) on South facing wall, two probes on both internal and external side,
h = 1.4 m and h = 2.1 m.

c
Position 3. On the Roof, on both internal and external side, in the roof center, about 1.5 m far from
the walls.

d
Position 4. Thermal flux probes positioned on the internal side of the roof and of the North-facing
wall, close to the superficial temperature probe h = 1.4 m.

e
Position 5. Downward oriented pyranometer outdoor positioned in the central point of each roof,
h = 50 cm.

f
Position 6. Energy meters measure the required electricity independently from their position. The
energy counter is positioned in the electric box of each roof, and the datalogger stations storage
these data.

g
The weather station is located over the roof of the department building, about 20 m far from the
rooms, h = 10 m.

The meteorological station (Fig. 2b), which is fully dedicated to this experiment, is
positioned over the roof of the university building, at about 10 meters far from the
Test-Room experimental field. This same equipment is integrated within the
system, therefore its data logger is connected to the indoor downloading position,
which is located into the university building as well.

The weather monitoring allows to describe the continuous climatological


conditions of the specific experimental site, in order to analyze the building
thermal-energy performance with respect to the dynamic climate boundary
conditions in terms of temperature trends, wind characterization, radiation
profiles, etc. [5]. Table 2 reports the description of the sensors and the monitored
parameters which constitute the overall weather station.

5. Results and discussion


This section deals with the analysis of the results of the continuous monitoring
concerning: the assessment of the indoor thermal behavior and the evaluation of
the Test-Rooms’ energy performance in winter conditions. The study is carried
out by taking into account the three main operating regimes of the Test-Rooms,
that are: (i) the operating system behavior, (ii) the transient regime, and (iii) the free
floating regime.

The comparative evaluation of the Test-Rooms’ hygrothermal performance is


carried out through the analysis of indoor air temperature, operative temperature,
mean radiant temperature and indoor humidity profiles of each Test-Room,
during the three investigated regimes. The thermal behavior of opaque envelope
components is also evaluated through the analysis of thermal flux through walls
and roof and their surface temperature. Additionally, the energy demand for
heating of each building is assessed.

5.1. Analysis of the performance in operating system conditions

5.1.1. Indoor thermal behavior


This section concerns the period November 13th 2012–January 6th 2013, while the
heat pump systems were continuously operating with constant air temperature
set-point corresponding to 20 °C. The three indoor thermal parameters vs. the
outdoor dry bulb temperature every ten minutes are reported in Fig. 3. The same
graphs show how the two Test-Rooms are characterized by equivalent indoor
thermal behavior, in particular for what concerns air temperature and operative
temperature. The weak difference is attributable to the different technical
characterization of the envelope components, described later.

Download : Download full-size image

Fig. 3. Comparison between the two Test-Rooms in terms of Tair, Top, Tmr vs. outdoor dry
bulb temperature.

In order to investigate the thermal behavior and to quantify the variability of the
considered thermal parameters, hourly and daily thermal profiles are now
assessed. In particular, Table 3 reports the maximum, minimum, and average
values of the same parameters registered during the operating system period, and
finally the coefficient of variation (VC), calculated through Eq. (1) where σ is the
standard deviation and T is the temperature:

(1)

Table 3. Main thermal parameters of the two Test-Rooms in operating system conditions.

Test-Room 1 Test-Room 2

Tair Tmr Top Tair Tmr Top

Maximum 25.94 32.78 28.38 25.73 33.00 28.78

Minimum 21.78 20.04 21.36 22.73 17.89 20.61

Average 23.56 21.96 22.76 23.96 20.57 22.27

VC 1.01 4.55 2.60 0.81 6.52 3.25

Table 3 shows that the difference between the average values of indoor air
temperature and operative temperature is less than 0.5 °C, while the main
differences are found in terms of mean radiant temperature. The reason why the
radiant properties of Test-Rooms’ envelopes present the largest singularities
should be due to the different construction technology of walls and roofs, and to
the different external reflective properties of the envelope elements. In particular,
the no-sloped roof is covered by bitumen-based membrane with very low
reflectance capability, which determines the roof overheating in sunny hours even
in winter conditions, and the relative mean radiant temperature increase of the
corresponding Test-Room (TR-2). In fact the same TR-2 presents a maximum level
of mean radiant temperature of 33 °C, with the highest variation coefficient
produced by the high daily thermal excursion, typical of solar absorptive
envelopes. Additionally, the presence of the air-gap layer in the roof of TR-1
determines a lower thermal flux entering the roof, which produces an overall
warmer operative indoor thermal condition than the other Test-Room. Focusing
on the operative temperature, as the fundamental parameter to investigate indoor
global thermal conditions [5], the difference between the average daily Top between
the two Test-Rooms is always less than 0.5 °C, and it is characterized by the same
consistent tendency along the entire period, which determines the possibility to
compare the two Test-Rooms also in terms of operative temperature, for future
coupled experiments. In fact, a very low variability of the standard deviation of the
difference between the operative temperature of TR-1 and TR-2 was found, i.e.
0.13 °C. The same conclusion is found through calculating the Pearson correlation
coefficient between the hourly values of operative temperature and mean radiant
temperature, by Eq. (2). These values, which correspond to 0.96 for both the
parameters, show very high strength of the positive linear relationship of the
hourly operative temperature and mean radiant temperature values between the
two Test-Rooms:

(2)

where Nm is the number of m pairs of measured data, TTR-1 and TTR-2 are the
monitored parameters in Test-Room 1 and Test-Room 2, respectively.

Lower linear correlation values of air temperature, which corresponds to 0.70 of


the Pearson coefficient r (2), are imputable imputable to the higher daily variability
of indoor air temperature measured in TR-2, which is produced by the continuous
on–off behavior of the heat pump system registered during the operating regime,
as shown in Fig. 4. The graph in Fig. 4 represents a typical daily thermal profile of
the operating system period (November 23rd, 2012) where, by comparing the two
Test-Rooms’ behavior in terms of temperature measured every 10 min, it is
possible to see that the Tair presents significant correlation during the course of
the day, consistently with the results of the entire period. In particular, Test-Room
2 is characterized by a slightly higher air temperature with higher variability, and
lower mean radiant temperature which produces a light difference in terms of
operative temperature, during the entire course of the continuous monitoring.
Therefore, Test-Room 1 is able to better preserve the indoor thermal stationary
conditions during the operative regime than Test-Room 2, and the internal
envelope surface temperature of Test-Room 2 is in general lower than in Test-
Room 1.

Download : Download full-size image

Fig. 4. Daily thermal profiles of the Test-Rooms during the operating systems’ regime.

5.1.2. Energy performance for heating


The comparative analysis of the two test buildings is carried out in terms of
energy requirement for heating during the same period November 13th, 2012–
January 6th, 2013. Fig. 5a reports the energy consumption vs. the outdoor dry bulb
temperature for both the Test-Rooms, measured every 10 min, where the high
correspondence is pretty evident. Focusing on one representative winter day of the
same period (December 21st, 2012), as reported in Fig. 5b, the same relationship is
even more evident. In fact, the two linear regression lines are basically coincident.
Additionally, the correlation coefficient r between the two Test-Rooms, calculated
through (2) between the hourly mean calculated energy values, corresponds to 0.96
while the average difference between the two Test-Rooms’ measurements is
0.06 Wh, with a daily standard deviation value of 1.2 Wh.

Download : Download full-size image

Fig. 5. Test-Rooms’ energy consumption for heating (measured every 10 min) vs. external dry
bulb temperature. Long term monitoring (a) and one day measurements registered on
December 21st, 2012 (b).

The observed high correspondence between the energy behavior of the two Test-
Rooms in winter conditions is described through a correlation coefficient of 0.98,
calculated between the daily energy requirement of the two Test-Rooms. These
results demonstrate that the two Test-Rooms, in these boundary conditions, could
be used to perform simultaneous continuous monitoring campaigns with the
purpose to investigate the different energy behavior of several building
optimization strategies, i.e. envelope innovative materials, glazing systems, HVAC
technologies.

5.2. Analysis of the performance during the transient period

5.2.1. Indoor thermal behavior


The analysis of the indoor thermal behavior in terms of air temperature, mean
radiant temperature and operative temperature was carried out for the twelve-day
period following the heat pumps’ switching off on January 7th, 2013. Fig. 6 reports
the temperature measurements every 10 min where the indoor transient cooling
path is progressively registered. An important correlation was found by comparing
the thermal trends of the two Test-Rooms. The average difference between the two
Test-Rooms in terms of these thermal parameters is always less than 0.43 °C, with
a standard deviation value equal to 0.19 °C for Tmr, Tair and Top as well.

Download : Download full-size image

Fig. 6. Progressive cooling trend of the two Test-Rooms during the transient regime in terms
of Tair, Top, and Tmr.

Daily average operative temperature is reported in Fig. 7, where the high standard
deviation values were produced by the progressive free-floating cooling of the
indoor environment of each Test-Room. All the analyses demonstrate that the two
Test-Rooms present equivalent behavior during the transient period, determined
also by a Pearson correlation coefficient equal to 0.99 for the three thermal
parameters taken into account, i.e. Tmr, Tair, Top. This finding was also confirmed
by each daily thermal profile, where Test-Room 1 describes equivalent path with
temperature profiles weakly higher than the Test-Room 2 of about 0.3 °C.

Download : Download full-size image

Fig. 7. Daily mean operative temperature values (±σ) of the two Test-Rooms during the
transient regime.

5.3. Analysis of the performance during the free-floating period

5.3.1. Indoor thermal behavior


The analysis of the thermal parameters after the transient period concerned the
15-day period between January 22nd and February 5th, 2013, when the heat pumps
were not working in both the Test-Room buildings, in order to monitor their free-
floating regime. Consistently with the previous analyses, the behavior of the two
Test-Rooms presented important similarities in terms of daily average Tair, Top,
and Tmr (Fig. 8). The registered values reported an average difference of 0.44 °C for
Tmr and 0.52 °C for Tair with the standard deviation values corresponding to 0.1 °C
for both the cases calculated on hourly database. Consistently with the previous
analysis, the thermal behavior of Test-Room 1 is weakly warmer than Test-Room 2,
with consistent Top difference of about 0.48 °C for the overall period.

Download : Download full-size image

Fig. 8. Daily mean temperature values (Tair, Top, Tmr) of the two Test-Rooms vs. external dry
bulb temperature in free-floating regime.

The analysis concerning each daily path shows that, after the transient progressive
cooling, the indoor temperature of both the Test-Rooms reached an overall
constant value lightly impacted by the daily variations of the external weather
conditions, both in terms of air temperature and mean radiant temperature.
Therefore, the buildings responded in equivalent way to the ante-meridiem
increase of outdoor air temperature and solar radiation, and to the following
afternoon cooling. The Top profiles show equivalent thermal behavior with TR-2
lightly colder than TR-1 (Fig. 9). Particular attention was paid to analyze the
weather conditions during the free-floating monitoring when, from January 29th
to February 2nd, warmer outdoor temperature was registered. In these days, the
two Test-Rooms registered higher differences due to differential reaction to the
external heating which produced higher temperatures of Test-Room 1. In the
same period, no evident correlation between external solar radiation and indoor
thermal behavior was registered.

Download : Download full-size image

Fig. 9. Comparison between the Top profiles of the two Test-Rooms with respect to outdoor
weather conditions in free-floating regime.

5.4. Analysis of envelope behavior


This section specifically concerns the analysis of the thermal behavior of roof and
external walls in relation to the thermal-energy behavior of the two Test-Rooms.

5.4.1. Roof and wall performance


Both external and internal surface temperature of the roof are evaluated in order
to investigate the in-field dynamics of different roof configurations, characterized
by the same calculated stationary properties. Fig. 10 reports the surface
temperature profiles and the thermal fluxes through the roof of each Test-Room
during the period November 13th, 2012–January 8th, 2013, when the heat pumps
were operating. Consistent thermal behavior is observed during the overall period,
when the internal surface temperature of the roof of TR-1 is higher than TR-2,
even if the opposite happens for the external surface temperature during the
sunny hours of the day. Therefore, the high solar absorbance of the waterproof
bitumen membrane in TR-2 is able to increase the daily thermal peaks of the
external surface of the roof, while during the night and during rainy or cloudy
days, the external surface temperature of TR-2 roof is not influenced by the
different optic-thermal behavior of the tiles with respect to the bitumen-based
membrane. As widely investigated [5–6; 14], this experiment confirms the
important dynamic effect produced by the solar reflectance capability of the
external layer of the roof, exposed to the solar radiation. Despite the winter period
is characterized by lower solar radiation, the thermal dynamic effect is clearly
identifiable in the roof performance. In fact, the optic-energy property of the
external surface of the roof, i.e. solar reflectance, is mainly responsible for the
higher external surface temperature of TR-2 than TR-1 by 11.0 °C and 2.3 °C in
terms of maximum daily peak and daily mean temperature, respectively.
Nevertheless, the internal roof surface temperatures of the Test-Rooms seem to be
not affected by the solar reflectance characteristic of the roof external skin. In fact,
the roof internal surface temperature in TR-1 is higher than in TR-2 by about 1.6–
1.9 °C, in terms of daily average values and daily peaks, respectively. This measured
surface temperature discrepancy, and the higher thermal flux registered through
the roof of TR-2 with respect to TR-1, contrast with the stationary calculated
values of thermal transmittance of the TRs’ roofs (Table 1).

Download : Download full-size image

Fig. 10. Comparison between thermal fluxes (a) and roof internal and external surface
temperatures (b) with respect to weather conditions.

In order to verify this element, the in-field thermal conductance of roofs is now
calculated. The analysis of in-field pure conduction behavior of walls and roofs is
here preferred, in order to avoid possible differential air temperature stratification
effects between the two Test-Rooms. Nevertheless, specific attention should be
paid to the real thermal contact resistance between adjacent layers of each wall and
roof, affecting in-field conductance measurements.

The measurements are taken into account for the overall period with operating
heating system [34]. Fig. 11a reports the converging to asymptotical values of
thermal conductance of the roofs, which are measured by following the procedure
reported in [35] for heavy elements, i.e. with specific heat per unit area of more
than 20 kJ/(m2 K), as described in Table 1.

Download : Download full-size image

Fig. 11. Roof (a) and wall (b) in-field measurement of thermal conductance: observed
convergence to an asymptotical value.

As expected, in the view of pure conduction, the in-field calculated value of roof
thermal conductance of TR-1 is much lower than the one calculated in TR-2, that
allows much higher thermal losses through the roof. This consideration is also
confirmed by the monitored thermal flux values during the same period, which
difference is 3.7 W/m2 on average in the controlled regime period. Additionally, as
already mentioned, the daily superficial overheating of TR-2 roof makes the daily
maximum peak grow up to 11 °C higher than TR-1 during the analyzed period,
while the TR-2 is colder in terms of minimum daily peaks by 0.4 °C. This singular
behavior determines a daily fluctuation of the external surface temperature by
21 °C in TR-2 and only 9 °C in TR-1, which produces important thermal stresses
for the TR-2 roof. Nevertheless, the internal surface of the roof is characterized by
consistent colder temperature in TR-2. These experimental results also showed
that the theoretical method commonly used to calculate roof thermal
transmittance [35] of multilayer roofs could produce non-negligible errors
specifically attributable to the in-field observed dynamic behavior of materials and
components. In this case, the presence of the air gap in the roof of TR-1
represents the main difference between the two roof structures of TR-1 and TR-2,
respectively. Additionally, this calculation method could underestimate the
dynamic effect of the position of the insulation layer within each multilayer
element, producing deviation between calculated and in-field measured values of
conductance.

5.4.2. Wall performance


The equivalent analysis, carried out for the walls of TR-1 and TR-2, shows a
consistent behavior with respect to the roof. In fact, the daily average fluctuation of
surface temperature of the North-facing wall in TR-2 (10.7 °C) is slightly higher
than in TR-1 (6.7 °C), and the internal surface temperature of TR-1 is consistently
higher than TR-2. The internal surface temperature of the North-facing wall of
TR-2, in terms of daily peaks and daily mean values, is lower than TR-1 by around
0.5 °C, showing a smaller but consistent overcooling effect compared to the roof.
The convergence trend of the conductance values to an asymptotical value are
reported in Fig. 11b. Also in this case, the calculated theoretical values
underestimates the observed thermal losses through the walls for TR-2 in
particular. This consideration is confirmed by the monitored thermal flux through
the walls, where the results are consistent with the conductance measurement. In
fact, the calculated thermal transmittance (surface-to-surface) of the wall is higher
by 31% and 46% with respect to the theoretical value for TR-1 and TR-2,
respectively. The corresponding thermal flux through the North-facing wall of TR-
2 is higher than the one of TR-1 by 1.2 W/m2 on average in the controlled regime
period. Possible motivation of such difference between theoretical and measured
thermal resistance is also imputable to the thermal contact resistances between
adjacent layers, affecting the total thermal resistance of each building multilayer
envelope. Therefore, the actual thermal resistance is different with respect to the
theoretical value. Additionally, in the case of North-facing walls, the solar
reflectance property of each wall does not differ evidently between each other,
since also the color of the painting in TR-2 wall was selected for its similar
properties with respect to the natural red brick finishing (Fig. 1). Additionally, the
North-facing wall, especially in winter conditions, is very poorly affected by solar
radiation effect. Therefore, the main gap between the walls of TR-1 and TR-2
could be attributable to the dynamic effect produced by the different position of
the insulation layer, determining deviation in the dynamic behavior of the wall.

5.4.3. Moisture and surface temperature analysis


The overall comparative analysis of the envelope between the two Test-Rooms
leads to several observations concerning relative humidity and temperature
dynamics of the two prototype buildings during the analyzed winter period. Fig. 12
reports the hourly profiles of internal surface temperatures of roofs, North-facing
and South-facing walls for each Test-Room, together with the indoor relative
humidity profile, both before and after the HVAC system turning-off (gray large
vertical line in the figure).

Download : Download full-size image

Fig. 12. Internal surface temperature profiles and indoor relative humidity behavior of TR-1
and TR-2.

The monitored data clearly show that the indoor relative humidity (RH) profiles of
TR-1 and TR-2 describe different behavior. In fact, the indoor RH of TR-2 slowly
grows during the operating HVAC system period and then it begins to grow faster,
after the HVAC system turning off. Therefore, the RH hourly profiles present large
differences by comparing TR-1 and TR-2 in transient and free floating period.
These periods are characterized by high RH values, which reach non-comfortable
ranges. Non evident Pearson correlation coefficients between RH and surface
temperatures are found [36] but the overall RH trend highlights significant
thermal phenomena. Internal surface temperature profiles of roof and walls in
each Test-Room show decreasing mutual difference with increasing relative
humidity trend. Surface temperature discrepancy between TR-1 and TR-2
envelope, after the HVAC turning-off, is less than 1 °C. Additionally, the roof
internal surface thermal behavior follows the same profile of the monitored walls
of the same Test-Room. On the other side, with lower RH values (before the HVAC
turning off ), the TR-1 roof internal surface tends to be warmer and the inner
ventilation layer of the roof, together with the more transpiring structure of the
same TR-1 roof, allows a better transpiration process. The same transpiration
capability, which is influenced by many characteristics of the structure and the
microclimate of the location [37], [38], is able to influence the envelope thermal
insulation behavior, with the following consequences in terms of in-field
measured conductance value. Given the same positioning and exposure of the two
prototype buildings, the same outdoor boundary climate conditions are assumed,
therefore the indoor moisture behavior is imputable to the envelope
configurations. The humidity perception is very different between TR-1 and TR-2,
where superficial condensation is observed on the lower part of the walls and of
the door. The non-transpiring insulation layer of TR-2 is positioned on the
external face of the walls, and the evaporation process is particularly difficult all
over the winter period. The high superficial condensation and humidity rate in
TR-2 makes walls and roof describe very similar internal surface thermal behavior,
much more than TR-1, where the roof is able to allow the evaporation process and
to be warmer and drier than walls. Therefore, the high moisture content in the
overall TR-2 envelope and in TR-1 walls produces increased in-field conductance
with respect to the theoretical values, as described in previous sections of the
work.

6. Conclusions
The research about building dynamic thermal-energy behavior mainly focused on
the analysis of summer performance of buildings, where dynamic boundary
conditions are properly considered as key variables determining building
performance. Nevertheless important thermal parameters affecting indoor
environment perception should be investigated even in winter conditions when, in
several national regulations, stationary hypothesis are commonly acknowledged to
be the only driving the design process and the analysis of the thermal-energy
performance of buildings. In order to verify this hypothesis, and to study dynamic
thermal-energy behavior of buildings in winter conditions, this paper dealt with
the study of a long-term continuous monitoring campaign of two prototype
buildings. These buildings presented the same stationary properties, as calculated
at the design stage, but they are characterized by different envelope materials and
technologies. Therefore, following Italian regulations, they should behave the
same in winter conditions. This paper discussed the validity of this simplified
hypothesis by taking into account several thermal-energy parameters, as
continuously monitored in winter 2012–2013.

To this aim, after presenting the detailed characterization of the monitoring


integrated technology and of the architecture of the Test-Room buildings, results
of experimental tests were analyzed to compare the two Test-Rooms’ behavior in
terms of indoor hygrothermal conditions and energy requirement for heating. The
results of this paper specifically concern the three-month winter continuous
monitoring campaign, when the free-floating regime, the transient and the
operative systems’ regimes were studied.

Results demonstrated that the two prototype full-scale buildings presented


equivalent long-term indoor thermal dynamics in terms of air temperature, while
larger difference in terms of mean radiant temperature was measured. In fact, the
average hourly air temperature difference for each monitoring period is never
higher than 0.5 °C, while mean radiant temperature difference is higher than 1 °C,
in different climate and operational conditions. Moreover, the Pearson correlation
coefficients between the indoor environment parameters of the two Test-Rooms
were very close to a unit value over the entire investigated period (always higher
than 0.9). Additionally, during the operating system continuous monitoring, the
heating requirement, registered every 10 min, confirmed the high correspondence
of the two Test-Rooms in the overall assessed winter period, when the Test-
Rooms’ energy consumption presented an average difference of 0.03 Wh.

Nevertheless, non-negligible mutual differences were found in terms of relative


humidity behavior of the Test-Rooms, for moisture content of the walls in
particular. The impact of the humidity content, particularly high in the less
transpiring building envelope (Test-Room 2), corresponded to the increase of the
in-field measured conductance of walls and roofs and the overall thermal flux
through the envelope with respect to the theoretically calculated values of
conductance. Therefore, the two buildings did not present the same thermal
transmittance value, as hypothesized and calculated at the preliminary design
stage. In particular, Test-Room 1 performed even better than expected only in the
case of the roof, while Test-Room 2 was found to increase thermal conductance of
both the wall and the roof. Registered thermal flux through the walls and the roofs
of the Test-Rooms were radically different (3.7 W/m2 and 1.2 W/m2 of difference,
on average for the roof and the wall, respectively). Additionally, the same Test-
Room (TR-2) presented higher air humidity level and lower mean radiant
temperature than TR-1.

Based on the results obtained in this study, the investigated realistic test-cells
representing Italian construction typicality, and having the same envelope thermal
stationary characteristics, were found to behave the same in terms of energy
requirement for heating, where the simplified stationary regime hypothesis was
mainly satisfied. Conversely, the thermal-energy dynamics of each Test-Room
presented several observed differences, primarily imputable to different external
surface solar reflectance characteristics of the roof and different transpiring
behavior of the envelope stratigraphy, e.g. position of the insulation layer. These
observed differences were able to impact the correspondence between predicted
and in-field measured thermal insulation capability of walls and roof, and the
overall indoor comfort perception, given the different local and global
hygrothermal behavior.

Acknowledgments
The authors’ acknowledgements are due to Cassa di Risparmio di Perugia
Foundation, for supporting the construction of the experimental field through the
BAIO (Italian acronym for Indoor–Outdoor Environmental Comfort) project.
Additionally, the authors would like to thank all the industrial partners of BAIO
project.

References
[1] G. Dall’O, L. Sarto
Potential and limits to improve energy efficiency in space heating in existing school
buildings in northern Italy
Energy and Buildings, 67 (2013), pp. 298-308
Article Download PDF View Record in Scopus Google Scholar

[2] U. Berardi
Clarifying the new interpretations of the concept of sustainable building
Sustainable Cities, 8 (2013), pp. 72-78
Article Download PDF View Record in Scopus Google Scholar

[3] F. Pernodet Chantrelle, H. Lahmidi, W. Keilholz, M. El Mankibi, P. Michel


Development of a multicriteria tool for optimizing the renovation of buildings
Applied Energy, 88 (2011), pp. 1386-1394
Google Scholar

[4] M. D’Orazio, C. Di Perna, E. Di Giuseppe


Green roof yearly performance: a case study of highly insulated building under
temperate climate
Energy and Buildings, 55 (2012), pp. 439-451
Article Download PDF View Record in Scopus Google Scholar

[5] A.L. Pisello, F. Cotana


The thermal effect of an innovative cool roof on residential buildings in Italy: results
from two years of continuous monitoring
Energy and Buildings, 69 (2013), pp. 154-164
Google Scholar

[6] E. Bozonnet, M. Doya, F. Allard


Cool roofs impact on building thermal response: a French case study
Energy and Buildings, 43 (2011), pp. 3006-3012
Article Download PDF View Record in Scopus Google Scholar

[7] Kolokotsa Da, C. Diakaki, S. Papantoniou, A. Vlissidis


Numerical and experimental analysis of cool roofs application on a laboratory
building in Iraklion, Crete, Greece
Energy and Buildings, 55 (2012), pp. 85-93
Google Scholar

[8] A.L. Pisello


Optic-energy performance of innovative and traditional materials for roof covering in
commercial buildings in central Italy
Advanced Materials Research, 884–885 (2014), pp. 685-688
View Record in Scopus Google Scholar

[9] G. Dall’O’, L. Sarto, N. Sanna, A. Martucci


Comparison between predicted and actual energy performance for summer cooling in
high-performance residential buildings in the Lombardy region (Italy)
Energy and Buildings, 54 (2012), pp. 234-242
Article Download PDF View Record in Scopus Google Scholar

[10] P. Wouters, L. Vandaele, P. Voit, N. Fisch


The use of outdoor test cells for thermal and solar building research within the
PASSYS Project
Building and Environment, 28 (2) (1993), pp. 107-113
Article Download PDF View Record in Scopus Google Scholar

[11] P.A. Strachan, P.H. Baker


Outdoor testing analysis and modelling of building components
Building and Environment, 43 (2008), pp. 127-128
Article Download PDF View Record in Scopus Google Scholar

[12] P.A. Strachan, L. Vandaele


Case studies of outdoor testing and analysis of building components
Building and Environment, 43 (2008), pp. 129-142
Article Download PDF View Record in Scopus Google Scholar

[13] P.H. Baker, H.A.L. van Dijk


PASLINK and dynamic outdoor testing of building components
Building and Environment, 43 (2008), pp. 143-151
Article Download PDF View Record in Scopus Google Scholar

[14] M.J. Jiménez, H. Madsen


Models for describing the thermal characteristics of building components
Building and Environment, 43 (2) (2008), pp. 152-162
Article Download PDF View Record in Scopus Google Scholar

[15] M.J. Jiménez, H. Madsen, K.K. Andersen


Identification of the main thermal characteristics of building components using
MATLAB
Building and Environment, 43 (2) (2008), pp. 170-180
Article Download PDF View Record in Scopus Google Scholar

[16] A. Androutsopoulos, J.J. Bloem, H.A.L. van Dijk, P.H. Baker


Comparison of user performance when applying system identification for assessment
of the energy performance of building components
Building and Environment, 43 (2) (2008), pp. 189-196
Article Download PDF View Record in Scopus Google Scholar

[17] M. Doya, E. Bozonnet, F. Allard


Experimental measurement of cool facades’ performance in a dense urban
environment
Energy and Buildings, 55 (2012), pp. 42-50
Article Download PDF View Record in Scopus Google Scholar

[18] M. Santamouris, N. Papanikolaou, I. Livada, I. Koronakis, C. Georgakis, A. Argiriou,


D.N. Assimakopoulos
On the impact of urban climate on the energy consumption of buildings
Solar Energy, 70 (2001), pp. 201-216
Article Download PDF View Record in Scopus Google Scholar

[19] A. Joudi, H. Svedung, M. Cehlin, M. Rönnelid


Reflective coatings for interior and exterior of buildings and improving thermal
performance
Applied Energy, 103 (2013), pp. 562-570
Article Download PDF View Record in Scopus Google Scholar

[20] C. Castellón, A. Castell, M. Medrano, I. Martorell, L.F. Cabeza


Experimental study of PCM inclusion in different building envelopes
Journal of Solar Energy Engineering, 131 (2009), pp. 041006-1-041006-7
Google Scholar

[21] X. Xu, J.E. Taylor, A.L. Pisello


Network synergy effect: establishing a synergy between building network and peer
network energy conservation effects
Energy and Buildings, 68 (2014), pp. 312-320
Article Download PDF CrossRef View Record in Scopus Google Scholar

[22] A. Bontemps, M. Ahmad, K. Johannès, H. Sallée


Experimental and modelling study of twin cells with latent heat storage walls
Energy and Buildings, 9 (2011), pp. 2456-2461
Article Download PDF View Record in Scopus Google Scholar

[23] S.P. Corgnati, E. Fabrizio, M. Filippi, V. Monetti


Reference buildings for cost optimal analysis: method of definition and application
Applied Energy, 102 (2013), pp. 983-993
Article Download PDF View Record in Scopus Google Scholar

[24] F. Rossi, A.L. Pisello, A. Nicolini, M. Filipponi, M. Palombo


Analysis of retro-reflective surfaces for urban heat island mitigation: a new analytical
model
Applied Energy, 114 (2014), pp. 621-631
Article Download PDF View Record in Scopus Google Scholar

[25] A.L. Pisello, M. Santamouris, F. Cotana


Active cool roof effect: impact of cool roofs on cooling system efficiency
Advances in Building Energy Research, 7 (2) (2013), pp. 209-221
CrossRef View Record in Scopus Google Scholar

[26] E. Bonamente, F. Rossi, V. Coccia, A.L. Pisello, A. Nicolini, B. Castellani, F. Cotana, M.


Filipponi, E. Morini, M. Santamouris
An energy-balanced analytic model for urban heat canyons: comparison with
experimental data
Advances in Building Energy Research, 7 (2) (2013), pp. 222-234
CrossRef View Record in Scopus Google Scholar

[27] A.L. Pisello, F. Cotana, L. Brinchi


On a cool coating for roof clay tiles: development of the prototype and thermal-
energy assessment
Energy Procedia, 45 (2014), pp. 453-462
Article Download PDF View Record in Scopus Google Scholar

[28] C. Buratti, E. Moretti


Experimental performance evaluation of aerogel glazing systems
Applied Energy, 97 (2012), pp. 430-437
Article Download PDF View Record in Scopus Google Scholar

[29] C. Buratti, E. Moretti


Glazing systems with silica aerogel for energy savings in buildings
Applied Energy, 98 (2012), pp. 396-403
Article Download PDF View Record in Scopus Google Scholar

[30] C. Buratti, P. Ricciardi, M. Vergoni


HVAC systems testing and check: a simplified model to predict thermal comfort
conditions in moderate environments
Applied Energy, 104 (2013), pp. 117-127
Article Download PDF View Record in Scopus Google Scholar

[31] F. Asdrubali, G. Baldinelli, F. Bianchi


Influence of cavities geometric and emissivity properties on the overall thermal
performance of aluminum frames for windows
Energy and Buildings, 60 (2013), pp. 298-309
Article Download PDF View Record in Scopus Google Scholar

[32] ISO 13786: 2007(E), Thermal performance of building components – dynamic thermal
characteristics – calculation methods.
Google Scholar

[33] M. Zinzi, E. Carnielo, S. Agnoli


Characterization and assessment of cool coloured solar protection devices for
Mediterranean residential buildings application
Energy and Buildings, 50 (2012), pp. 111-119
Article Download PDF View Record in Scopus Google Scholar

[34] G. Desogus, S. Mura, R. Ricciu


Comparing different approaches to in situ measurement of building components
thermal resistance
Energy and Buildings, 43 (2011), pp. 2613-2620
Article Download PDF View Record in Scopus Google Scholar

[35] EN ISO 6946: Building components and building elements. Thermal resistance and
thermal transmittance. Calculation method. (EN ISO 6946:2007).
Google Scholar

[36] J.L. Nguyen, J. Schwartz, D.W. Dockery


The relationship between indoor and outdoor temperature, apparent temperature,
relative humidity, and absolute humidity
Indoor Air (2013), 10.1111/ina.12052
(Epub ahead of print)
Google Scholar

[37] D. D’Agostino
Moisture dynamics in an historical masonry structure: the Cathedral of Lecce (South
Italy)
Building and Environment, 63 (2013), pp. 122-133
Article Download PDF View Record in Scopus Google Scholar

[38] F. Ascione, N. Bianco, R.F. De Masi, F. De’Rossi, G.P. Vanoli


Simplified state space representation for evaluating thermal bridges in building:
modelling, application and validation of a methodology
Applied Thermal Engineering, 61 (2013), pp. 344-354
Article Download PDF View Record in Scopus Google Scholar

View Abstract

Copyright © 2014 Elsevier B.V. All rights reserved.

About ScienceDirect Remote access Shopping cart Advertise Contact and support Terms and conditions Privacy policy

We use cookies to help provide and enhance our service and tailor content and ads. By continuing you agree to the use of cookies.
Copyright © 2021 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V.
ScienceDirect ® is a registered trademark of Elsevier B.V.

You might also like