Sei sulla pagina 1di 41

Challenges in Build-Back-Better Housing Reconstruction Programs for

Coastal Disaster Management: Case of Tacloban City, Philippines

James Michael Ong1, Ma. Laurice Jamero2, Miguel Esteban3, Riki Honda4 and Motoharu Onuki5

1
Graduate Student, Graduate Program in Sustainability Science, The University of Tokyo

Mailing Address: Rm 334, Building of Environmental Studies, 5-1-5 Kashiwanoha, Kashiwa City,

Chiba 277-8563 JAPAN

E-mail: ong@sustainability.k.u-tokyo.ac.jp, ong.james.michael@gmail.com

2
Graduate Student, Graduate Program in Sustainability Science, The University of Tokyo

E-mail: laujamero@gmail.com

3
Project Associate Professor, Graduate Program in Sustainability Science, The University of Tokyo

E-mail: esteban.fagan@gmail.com

4
Professor, Graduate Program in Sustainability Science, The University of Tokyo

E-mail: rhonda@k.u-tokyo.ac.jp

5
Professor, Graduate Program in Sustainability Science, The University of Tokyo

E-mail: onuki@k.u-tokyo.ac.jp

1
Abstract

In November 2013, super-typhoon Haiyan left more than 53,000 damaged houses in Tacloban

City. The Philippine government organized housing reconstruction programs relating to on-

site reconstruction and off-site relocation. However, these programs are faced with complex

implementation issues.

The present study investigated the mechanisms of three types of housing reconstruction

programs (i.e. owner-driven on-site reconstruction, community-driven off-site relocation and

contractor-driven off-site relocation cases) and discussed the sustainability challenges by

analyzing the gaps between community needs and the program outputs. Key informant

interviews with government officials and non-government organization representatives, and

semi-structured questionnaire surveys with beneficiary households were conducted in March

2015.

The results showed that on-site reconstruction was delayed due to insufficient and poorly

implemented assistance schemes relating to reconstruction (e.g. materials, skills training),

while offsite relocation was delayed by prolonged land acquisition and issues in

subcontracting. Disruption of critical infrastructure such as the water utility service and lack

of livelihood significantly affected the level of satisfaction of respondents with the recovery

progress. The study also found that the no-dwelling-zone policy was not strictly enforced as

houses were still being rebuilt in high risk areas near the coast. Lastly, permanently relocated

residents struggle, especially with regard to their source of livelihood, as relocation sites are

inaccessible and located away from employment opportunities.

Keywords: Multi-Layer Safety System, Coastal Land Use Policy, Shelter, Resilience, Sustainable

Development, No Dwelling Zone

2
Introduction

Typhoon Haiyan (Local Name: Yolanda) struck the Philippines on November 8, 2013 at almost the

peak of its power, causing enormous damage to Leyte, Samar and many other islands in the Visayas

region, the group of islands in Central Philippines. The maximum sustained wind speeds were around

160 knots, some of the largest in recorded history [Schiermeier, 2013; Takagi et al., 2015]. The strong

winds, together with the typhoons extremely low central pressure (895 hPa) caused great damage to

housing, infrastructure and vegetation, leaving behind bare mountains and flattened fields. All

informal dwellings were torn apart and even well-built official government buildings and schools

suffered serious damage, with their roofs being blown away and most windows shattered [ Takagi et

al., 2014]. A large storm surge was also generated by the typhoon, which engulfed several coastal

towns and caused large damage to Tacloban City and the coastline around it [Takagi et al., 2014;

Tajima et al., 2014]. 6,245 individuals were reported dead, 28,626 injured and 1039 are still missing

[NDRRMC, as of 6 March 2014]. It was one of the deadliest disasters to have affected the country,

surpassing that of the 1991 floods in the Ormoc region in western Leyte, where 5,101 perished due to

Tropical Storm Thelma. The number of damaged houses was estimated to be 1,140,332, with 550,928

of them being completely destroyed [NDRRMC, 2014]. The total economic loss could be over 35,000

million pesos (776 million USD), probably the most expensive natural disaster in the history of the

country [Brown, 2013], despite the fact that it frequently suffers heavy loses to agriculture and

housing due to these weather systems [Stromberg et al., 2011; Esteban et al., 2012].

The exceptionally large storm surge was responsible for many of the deaths. A maximum inundation

height of 6-7 m. was observed in Tacloban city, where the largest number of casualties took place

[Shibayama et al., 2014; Tajima et al., 2014; Lagmay et al., 2015; Takagi et al., 2015]. The storm

surge was rather unusual in height and characteristics, possibly due to local amplification of the water

surface elevation due to seiche effects inside Leyte Gulf [Mori et al., 2014] and as a result of various

bathymetric features in the area, which made it manifest itself as a Tsunami-Like Surf Beat [Bricker

and Roeber 2015]. However, the large death toll was caused not only due to large size of the storm

surge but also due to issues related to the level of knowledge and awareness by local residents on
3
what is a storm surge [Esteban et al., 2014; Esteban et al., 2015; Lagmay, 2015; Leelawat et al.,

2014].

Typhoon disaster recovery processes are normally divided into rehabilitation and reconstruction

phases. This is often a complex period, when many difficult decisions must be taken, and the

understanding of recovery may differ between people who use this term and fluctuate depending on

the situations in which it is used [Matsumaru, 2015]. The rehabilitation phase typically represents a

quick repair of infrastructure and facilities to restore the economic and social functions of disaster-

affected areas. This might take several weeks to more than a year, depending on the extent of the

damage and infrastructure to be repaired [Matsumaru, 2015]. After the rehabilitation is over, the

reconstruction phase is seen as a long-term restoration that includes not only physical improvement of

the affected communities but also the revival of livelihoods, economy and industry, culture and

traditions and the environment [Matsumaru, 2015; Esteban et al, 2015].

Recent perspectives towards post-disaster reconstruction aim to enhance resilience in communities

against future disasters. In this paper, resilience is defined as the systems capacity to absorb

disturbance, in this case, storm surge as the coastal hazard, and re-organize into a fully functioning

system. It involves not only the systems capacity to return to its original state, but also to move to a

more advanced state through learning and adaptation [Adger et al. 2005; Klein et al., 2003; Folke,

2006; Cutter et al. 2008]. In order to achieve these more advanced resilient states, the concepts of (1)

multi-level safety system, (2) build-back better principle and (3) sustainable socio-economic

development has to be integrated into the resilience framework.

Within such a multi-layer safety system, three safety layers can be distinguished [National Water Plan

of the Netherlands, 2012; Esteban et al., 2015; Esteban et al., 2013; Tsimopoulou et al., 2012;

Tsimopolou et al., 2013]:

Layer 1 - Prevention: encompasses various measures such as breakwaters or dykes that are aimed

at preventing seawater from inundating areas that are usually dry

4
Layer 2 - Spatial Solutions: using spatial planning and adaptation of buildings to decrease the loss

if a flood does occur. The spatial arrangements that are part of layer 2 measures include the

relocation of settlements away from areas at greatest danger of flooding, for example.

Layer 3 - Emergency Management: includes organizational preparation for floods such as disaster

plans, risk maps, early-warning systems, evacuation and medical help. The main focus of layer 3

measures is the reduction of risk to human life.

While traditional coastal disaster management in developed countries still tends to focus on

implementing structural measures during post-disaster reconstruction, current sustainability science

studies indicate the need to integrate non-structural (including mostly layer 2 and 3) measures (e.g.

coastal management policies and urban planning) to improve community resilience against future

events. This is even more important for the case of developing countries such as the Philippines,

whose resources for constructing infrastructure are often limited. The present paper will particularly

focus on the development of layer 2 strategies, which require the relocation of those living in areas at

risk into permanent housing elsewhere.

Moreover, the Build Back Better principle is very important in the attempt to create more resilient

societies by reducing the vulnerability to future risks [Schilderman, 2010]. This entails the

implementation of disaster risk reduction measures with an optimum combination of structural and

non-structural measures in the reconstruction process. Different approaches are required for Build

Back Better reconstruction, depending on the type of hazard. Since the damage by storm surges is

mainly caused by water, topographic conditions strongly influence damage patterns. Water intrusion

can be controlled by constructing structural countermeasures, which have been identified as having

future potential for reducing disaster risk, although these require financial investments that would

appear too large for Tacloban City [City Government of Tacloban, 2014]. The preservation of human

lives can be achieved if adequate evacuation systems are in place, as typhoons can be predicted days

in advance. Thus, for the case of Tacloban one of the key components of any Build Back Better plan

5
entails the relocation of coastal communities to safer areas to avoid the recurrence of future disasters,

which can be defined as a layer 2 countermeasure.

Taking the perspective that the post-disaster recovery phase can be a window of opportunity for

disaster risk reduction [Paul, 2011], disasters are now also being considered as a development

opportunity [Asgary et al., 2006, Harrington, 2005, Wisner et al., 2004, Thiruppugazh, 2007,

Palliyaguru and Amaratunga, 2011]. In this regard, reconstruction projects should foster sustainable

development, not only addressing the reconstruction of physical structures, but also provide new

employment opportunities, improve the quality of life, maintain resource equity and service

distribution to the affected communities [Palliyaguru and Amaratunga, 2011]. The success of recovery

at the household and community level depends on several factors. At the household-level, families

recover at different rates due to the influence of factors such as socio-economic and demographic

characteristics of the individuals, the amount of resources households own or the amount of financial

assistance they receive from external sources [Paul, 2011]. Meanwhile, at the community-level, the

residential, commercial, industrial, social and lifeline components need to be adequately addressed

[Mileti, 1999]. Delay of recovery can also occur due to the setting of new codes or policies and

funding programs on the building process, damage of the structures, horizontal and vertical

integration or social participation into local, regional and national networks [Tobin and Montz, 1997],

prior disaster experience and effective leadership [Paul, 2011]. Hence, to ensure faster recovery,

mapping out an inclusive and holistic approach to address the needs of the disaster-affected

communities is important. Although addressing the socio-economic needs pose challenges, these can

be converted into development opportunities [UNDP, 2006].

The present study investigates the mechanisms of three different housing reconstruction programs. It

also discusses the sustainability challenges in these programs, by analyzing the gaps between

community needs and the program outputs. Among the phases in disaster management, the recovery

process, which involves the reconstruction phase, is typically the least understood due to its complex

dimensions [Berke et al. 1993; Smith and Wenger, 2006]. More importantly, this study compares the

6
different approaches employed to the housing reconstruction programs in terms of the satisfaction of

the beneficiaries, which is rarely observed in literature. Hence, this paper aims to contribute to the

pool of literature regarding the challenges to these housing relocation projects as a layer 2 spatial

solution against future coastal disasters. Such attempts at relocation can easily fail if the new location

is inconvenient or does not offer enough livelihood options [Suzuki, 2012], requiring careful analysis

to ensure the success of the projects.

This paper is composed of 7 sections. Section 2 discusses the changes in the coastal land use policy of

Tacloban after typhoon Haiyan. Based on this policy, Section 3 explains about the various shelter

options that were made available to households who needed to be relocated from their original

locations. Next, section 4 introduces the three selected case study sites and the methodology used for

investigating them, while Section 5 outlines the displacement pattern of affected households from

their original locations based on the survey and key informant interviews. Then, Section 6 focuses on

identifying the project management issues encountered in the housing reconstruction program.

Through a more holistic approach, Section 7 examines the challenges of linking the programs to

sustainable socio-economic development as a means of ensuring successful relocation. Finally,

Section 7 summarizes lessons that can be learned from Taclobans housing reconstruction programs

and implementation experience.

7
Coastal Land Use Policy

Prior to the disaster, there was an inadequacy in terms of layer 2 countermeasures in Tacloban.

According to the local government agency (Tacloban City Housing Office), the National Government

secured ownership of the citys coastline within 20 m. from high tide going inland to control

development and therefore, ban construction of houses (Esteban et al., 2015b). However, several

informal settlements had been built up right up to the coastline and even extended onto the sea. These

settlements were particularly badly hit, with all wooden constructions swept away by the storm surge.

As a consequence of the event, authorities have revised the coastal land use policy of Tacloban, which

is now being implemented during the recovery phase [City Government of Tacloban, 2014].

Table 1 shows the development of the coastal land use policy of Tacloban. While the policy initially

designated a No Build Zone (NBZ) along the citys coastline in the earlier stages of recovery, it was

later revised to distinguish between No Dwelling Zone, Unsafe Zone, and Safe Zone. [ Presidential

Assistance for Rehabilitation and Recovery (PARR), 2014].

As of April 2014, the No Dwelling Zone follows a 40 m. limit that has been arbitrarily set by the

National Government until the in-depth risk assessment is finished and formal hazard maps are

finalized. The National Government initially cited the provision from Article 51 of the Water Code of

the Philippines as basis for the 40 m. buffer distance [Philippine National Government-DENR, 1976;

Philippines Shelter Cluster, 2014a]. The limit has consequently been questioned by NGOs as the Code

only protects water sources and does not primarily ensure public safety. Nonetheless, the National

Government reasons that this limit is a conscious effort to protect its local coastal communities

against future disasters.

8
Based on the recent rehabilitation plan of Tacloban City, the No Dwelling Zone, which is now to be

designated through the hazard maps released by the national government agencies [Department of

Environment and Natural Resources (DENR) et al., November 5, 2014; Lapidez et al., 2014], would

ban all houses, hotels or hospitals from the area, though it would allow some tourism, port and

recreational activities to take place. Furthermore, the construction of any buildings (i.e. commercial,

residential or industrial) in land elevation under 5 m. from the high water mark would be restricted to

low density and low rise developments [City Government of Tacloban, 2014]. Coastal zones along the

bays and San Juanico strait would be designated as recreation zones and replanted with mangroves for

protection as a resilience strategy.

For all this to happen it would be necessary for those presently living in areas close to the water to be

relocated, and a number of permanent relocation projects are currently underway, with a target of

more than 10,000 new houses to be built [City Government of Tacloban, 2014]. Such reconstruction

would not only remove people away from danger, but the quality of the houses that are being built

(largely made of concrete with a steel roof) is generally superior to the wooden houses typically

present in informal settlements (according to the concept of Build Back Better).

Lastly, it is important to note that this paper adopted the 40-meter buffer distance set by the National

Government in the time of transition to recovery phase as the definition of No Dwelling Zone (NDZ).

9
Shelter Options in Tacloban City

After Haiyan, government and non-government organizations provided various shelter options to

affected communities in stages, depending on whether or not they were originally located in safe

zones or no dwelling zones [DSWD et al. 2014] These options included tents (Figure 1a, short-

term) and evacuation centers (Figure 1b, short-term), bunkhouses Figure 1c, medium-term) and

transitional shelters (Figure 1d, medium-term) and permanent housing in the original housing location

(Figure 1e, long-term) or relocation sites (Figure 1f and 1g, long-term), as shown on Table 2 [City

Government of Tacloban, 2014]. Like bunkhouses, transitional shelters also accommodate households

from no dwelling zones awaiting permanent relocation. However, while bunkhouses are made up of

wooden row houses, transitional shelters consist of single, detached native houses. Transitional

shelters are also often situated within close proximity of the permanent houses to which residents

would be relocated. Most affected households moved into these bunkhouses and transitional shelters,

although some families opted to temporarily move in with their relatives instead.

Permanent relocation is only offered to households originally living in no dwelling zones. As there

are numerous government agencies and NGOs offering various forms of housing assistance, the

Tacloban City Housing Office acted as a coordinator during beneficiary selection to prevent the

duplication of efforts. Still, these agencies and organizations ultimately applied their own processes

and set of criteria in choosing target beneficiaries based on their respective program objectives.

Processes usually included beneficiaries writing letters to NGOs and/or drawing by lots, or local

officials endorsing a list of households in need of housing assistance to agencies and organizations.

On the other hand, while selection criteria varied to a certain extent, vulnerability was often

10
considered, thus giving preference to families with elderly, pregnant women, lactating mothers and

children members.

In terms of housing design, government agencies and NGOs both complied with the revised minimum

housing design standards set by the Department of Public Works and Highways (DPWH) after

Haiyan. To ensure that permanent housing designs are typhoon-resilient, DPWH now requires one-

storey infrastructures to withstand a wind load design criterion of 250 kph [Regala, 2014]; note that

Haiyans maximum windspeed was around 300 kph. As a result, all permanent houses are now

concrete structures with steel roofing. Moreover, in line with the Build-Back-Better principle, the

standard housing design recommends building on stronger foundations and increasing the size of

structural elements reinforcements, beam and columns- with better connection details [Regala,

2014; Philippines Shelter Cluster, 2014b].

After preparing the reconstruction plans and housing designs, national government agencies and

NGOs again coordinate with the City Housing Office (local government agency), and seek their help

in securing necessary permits during the pre-construction phase. The total combined target number of

housing units is 14,433, although only a fraction has been completed (123 families have moved in as

of March 2015). At the same time, 1,027 and 627 families are still residing in bunkhouses and

transitional shelters, respectively.

11
Case Study Sites and Methodology

In order to understand the relocation process at the various study sites the authors conducted

structured questionnaire surveys in March 2015 with local residents and displaced persons via a

cluster sampling method, with the community leaders acting as key resource persons in the selection

of respondents. Clusters were formed from areas with different types of housing reconstruction

approaches in Tacloban.

The surveys were conducted in three different locations, as shown on Figure 2. At each location the

aim was to capture the respondents housing reconstruction experiences, including a) their needs

versus assistances received from government and private sources, b) the factors that influenced their

decision to relocate (or not to relocate), and c) their involvement or participation in the reconstruction

process.

The survey was further complemented by focus group discussions with beneficiary households

(between 5 to 8 beneficiary households in one round of discussion for each project site visited), and

key informant interviews with government and non-government organizations (see Table 3).

Discussions and interviews highlighted the challenges faced by households now residing in permanent

housing locations as well as the various initiatives they have launched for coping with various

problems.

12
Finally, survey data was analyzed statistically via significance testing (Wilcoxon-Mann-Whitney Test

for categorical variables, Spearmans Rank Correlation for ordinal variables and One-way Analysis of

Variance for comparison of groups). The results are presented in the next section.

4.1. Magallanes Avenue: Owner-driven On-site Reconstruction (Case 1), (n 1=53)

The Magallanes area is composed of mostly informal settlements built along the coastline of

Barangay 52, 54 and 57 in the South Coast of Tacloban with around 750 families in total (see Figure

2). As most of the community is now classified by the National Government as a no dwelling zone,

rebuilding of houses is discouraged in the area. However, many residents still opted to do so and

returned from temporary shelters to their original locations to rebuild their wooden homes on a self-

help basis. Hence, the authors referred to this as owner-driven relocation, as the owners are

essentially in control over the rebuilding process of their houses where they were originally located

(on-site) by using their own means or with external financial and technical assistances.

4.2. GMA Kapuso Foundation Housing: Community-driven (Participatory) Off-Site Relocation

(Case 2), (n2=52)

Global Media Arts (GMA) Kapuso Foundation Housing is a permanent relocation site being

constructed in Barangay 105 in the North coast of Tacloban to accommodate beneficiary households

previously residing in no dwelling zones. GMAs residents originally come from the coastal

barangay of San Jose (Barangay 88), about 24 km to the south (see Figure 2). They are usually

families with 7 members and above, in line with GMAs specific criteria for beneficiary selection.

GMA Kapuso Foundation Housing features concrete, row houses with a floor area of 42 sq. m. mainly

constructed by private contractors with funding from one of the country's largest TV networks, GMA.

Nonetheless, beneficiary households also participate in project implementation via a sweat equity

agreement with GMA that stipulates 500 hours of construction work. The authors thus termed this
13
community-driven relocation as the entire community is highly involved in the project of relocating

to an area away from where they used to reside (off-site).

Aside from the actual housing units, a school with 20 classrooms is also included in the plan.

However, as of March 2015, most of the houses were still under construction and only 106 out of 400

target number of units have been completed and occupied.

4.3. NHA Ridgeview Socialized Housing: Contractor-driven off-site relocation (Case 3), (n 3=12)

Ridgeview is also a permanent relocation site located in Barangay 97 in the Mid-coast. It is one of the

National Housing Authoritys (NHA) 13 project sites for Tacloban city residents. This case was

termed contractor-driven as the NHA has engaged with private contractors for the implementation

of its concrete, loftable, row housing design across all project sites. Each unit costs PhP 295,000 or

US$6,500, has a floor area of 22 sq. meters (another 11 sq. m. is loftable) and is built to withstand 250

kph wind speed. However, as of March 2015, only 17 out of 1,000 target units had been completed

and occupied at NHA-Ridgeview. Residents also originally come from San Jose (Barangay 88) which

is about 21 km south of the site (see Figure 2).

14
Displacement Patterns

5.1. Location of Original Residence of Respondents

Figure 3 presents the distribution of the respondent samples according to the location of their original

houses, either inside or outside the No Dwelling Zone (NDZ). Across all study sites most

interviewed households lived inside NDZs (i.e. 74%, 75% and 100%, respectively). This was

expected of Cases 2 and 3, which are both off-site permanent relocation programs, as beneficiary

selection criteria prioritized precisely those living in the NDZ. However, the same finding raises

concerns for Case 1 (on-site reconstruction) as it confirms that most residents have indeed rebuilt in

unsafe areas. Analysis in the succeeding sections further stratified the respondents from Case 1 (on-

site reconstruction) into those living inside the NDZ (Case 1a) and outside the NDZ (Case 1b) to

provide a clearer understanding of the responses in the survey area.

5.2. Timeline of Relocation

Through the questionnaire surveys, it was possible to ascertain the movement of respondents from

their original houses into evacuation centers or tents (short-term), bunkhouses or transitional shelters

(medium-term) and permanent relocation sites (long-term), as shown in Figures 4, 5 and 6.

At the onset of the super-typhoon in November 2013, respondents from Case 1 (on-site

reconstruction, n=53) transferred from their own houses to evacuation centers or tents (92%) or to

15
their relatives houses (8%). However, it is important to note that there is uncertainty about the exact

timing of this transfer due to data limitations during the immediate disaster relief period (shown as a

grey-shaded region in Figure 4). In February 2014, about three months after the event, the households

then started to leave the short-term shelter options to return and rebuild their houses at their original

locations, despite these now being part of the NDZ.

Residents now living in Case 2 (community-driven off-site relocation) also left evacuation centers,

tents or their relatives houses after 3 months. However, rather than going back to their original

housing locations, they moved in to transitional shelters or bunkhouses instead (60%), where they

stayed for 9 months. Relocation to the North Coast, which is 24 km away from their original housing

location, began around August 2014 once the first permanent houses were completed.

Lastly, for Case 3 (contractor-driven off-site relocation case), beneficiaries only began to transfer to

the housing units around November 2014, a year after Typhoon Haiyan struck. As of March 2015,

only 17 out of the planned 1000 units have been completed and occupied, signaling delays in housing

construction and overall recovery process as compared to Case 2 (community-driven relocation site),

where 106 out of 400 units (27%) have already been completed.

It can be noted that Case 2 residents started to move in to their housing units in August 2014, 3

months earlier than Case 3. However, in terms of moving in to transitional shelters, most Case 3

residents were able to move in earlier (by May 2014) compared to Case 2 as some opted to stay in

tents, their own houses or their relatives houses instead.

16
Project Management Challenges to Post-disaster Housing Programs

The housing projects in Tacloban, including both on-site reconstruction and off-site relocation

programs, are encountering problems in the different stages (i.e. preparation, design and construction

stages) of project planning and implementation. These interconnected challenges are highlighted in

this section.

17
6.1 On-site Reconstruction Inside and Outside the No Dwelling Zone (NDZ)

Households located in areas highly affected by storm surge were provided with various schemes [City

Government of Tacloban, 2014] such as:

assistance for self-recovery offered by government and non-governmental organizations in the

form of shelter repair kits,

socialized housing program or permanent relocation to sites in the North Coast of Tacloban,

and

Community Mortgage Programs (CMP), where an organized group of beneficiaries can

purchase lots through long-term loans at socialized rates.

The first option was made available to residents living outside the NDZ. In addition, the government

prescribed guidelines on how to build safer houses in safe zones [Philippines Shelter Cluster, 2014b].

However, some rebuilt houses in Case 1b (on-site reconstruction, outside the NDZ) were observed to

have not followed this accordingly.

On the other hand, only the second and third options were made available to residents living inside the

NDZ as the revised coastal land use policy of the city government requires them to relocate to areas

with lower risk to storm surge instead. However, as discussed in Section 5.2, several residents of Case

1a (on-site reconstruction inside the NDZ) still rebuilt inside the NDZ, indicating poor

implementation of the policy. This situation also points to problems regarding the availability of

assistance to all households that need to be relocated, delay in the construction of the resettlement

houses, and lack of information dissemination on assistance options to target beneficiaries. According

to the Tacloban City Housing Office (local government agency), the 14,433 target number of units

only initially accommodates 36 barangays located within NDZ and is inadequate to serve all the

families needed to be displaced to a safer zone.

18
6.2 Off-site Relocation

The main challenge to project implementation in Cases 2 and 3 for off-site permanent relocation is the

delay in housing construction. As of March 2015, only 123 out of 14,433 target housing units have

been completed and occupied across all government- and non-government-organization-funded

projects.

On the one hand, land acquisition proved to be challenging and time-consuming as reaching an

agreement with land owners needed thorough negotiation. On the other hand, especially for Case 3

(contractor-driven off-site relocation), subcontracting has become a serious issue as many contractors

have been distributing part of their obligation to subcontractors, compromising construction quality

and materials. Interview with Tacloban City Housing Office revealed that a number of materials have

already been returned to suppliers due to poor quality, and an NGO-funded (Habitat-for-Humanity)

housing reconstruction project has already been stopped due to the use of sub-standard materials.

Such poor construction management may only inadvertently increase risks to future disasters.

19
Linking Housing Reconstruction Projects to Sustainable Development

The new paradigm for post-disaster reconstruction recognized that disaster can be an opportunity to

build back better rather than merely satisfying the demand of communities to return to their original

state immediately [Kates et al., 2006; Thiruppugazh, 2007; World Bank, 2014]. Hence, decision-

makers have a choice of whether to restore the status quo or to enhance development in disaster-

stricken regions. Following the above section on challenges to housing reconstruction projects, this

section highlights the challenge to incorporate soft solutions into these infrastructure-based projects in

order to promote a more holistic approach to recovery. The analysis below evaluates the need for this

holistic approach by checking the level of satisfaction of households with the housing reconstruction

programs (i.e. whether or not the assistance received was related to their actual needs). It also

identifies the types of assistances that would lead to greater satisfaction for each case of housing

reconstruction project.

Respondents from all cases received a variety of assistance types, as shown in Figure 7. Financial aid

was given by the national government (Department of Social Welfare and Development via 4P

Poverty Alleviation Program) and an international NGO (Tzu Chi Foundation via outright donation or

cash-for-work). Construction materials or shelter repair kits (containing plywood, lumber, galvanized

iron sheet, hammer, handsaw and nails approximately worth around US$ 400) were also provided by

other humanitarian agencies (International Organization of Migrants, Oxfam International and Red

Cross). Finally, however, as not everyone can be relocated at the same time, only households meeting

a set of criteria were selected by government agencies and NGOs as priority for permanent relocation.

7.1 Level of Beneficiary Satisfaction and Project Performance

20
Household respondents were asked to indicate their level of satisfaction regarding their post-disaster

status, which also reflects their satisfaction with assistance received from government agencies and

NGOs. This parameter had already been used in past literature to measure project performance

[Lizarralde, 2009; Karunasena and Rameezdeen, 2010] and user and community participation

[Barenstein, 2008; Bouraoui and Lizarralde, 2013; ESSC, 2014]. In this paper, the level of satisfaction

was tested for significance against potential variables coming from 6 main categories including

financial aid, safety of location, utility service, community initiatives, livelihood opportunities and

type of assistances received. Finally, the significant variables were further assessed as to how these

parameters could bring future challenges to the communities.

Figure 8 shows the level of satisfaction of the households for the different case types; note that lower

values indicate higher satisfaction level based on a 5-point Likert scale. Case 2 (community-driven

off-site relocation) had high level of satisfaction (mean= 1.60), while Cases 1 (on-site reconstruction)

and 3 (contractor-driven off-site relocation) had relatively neutral satisfaction (means= 2.54 and 2.64)

and slight level of dissatisfaction (mean= 3.67), respectively.. Comparing the means of satisfaction

level via one-way analysis of variance, authors found a statistically significant difference (F=25.94,

p<0.001) amongst the three cases. Hence, further analysis was done to better understand why such

difference in satisfaction level arose.

Significance testing was employed to assess which variables significantly affected the level of

satisfaction of the households surveyed. Statistical parameters representing the degree (effect size)

and direction of relationship and significance level are indicated in Table 4.

21
For Case 1a (on-site reconstruction, inside the NDZ), shelter repair kits significantly affected the level

of satisfaction of the respondents. Generally, the main objective of distributing shelter repair kits is to

promote self-recovery, particularly in terms of shelter. However, as the repair kits were distributed to

residents living inside the NDZ, it enabled them to rebuild in the high-risk zone as well. This situation

could have been due to misallocation of aid as well as miscommunication of the NDZ policy to

affected households. Based on Figure 6, approximately 32% of the respondents of Case 1a were able

to receive this assistance. Significance testing also found livelihood training assistance as a significant

variable for determining the satisfaction level. Based on Figure 6, only less than 10% of respondents

received such type of aid, potentially explaining their low satisfaction level. Hence, their perceived

level of satisfaction on their recovery status depends on the types of assistances received and how

successful these assistances were given to specific target households in the communities, as similarly

emphasized by Coppola [2007].

On the other hand, for Case 1b (on-site reconstruction, outside the NDZ), the variables found to

significantly influence the level of satisfaction included estimated housing damage and water utility

disruption duration. While all housing structures inside the NDZ were completely washed away by the

storm surge, some structures outside the NDZ (30%) were spared and only suffered slight to partial

damage (Figure 9), contributing to a higher level of satisfaction. Another important factor for

satisfaction is water availability, indicating the need to prioritize the restoration of access to water

in the post-disaster reconstruction phase [Palliyaguru and Amaratunga, 2011]. Utility

services, including water and electricity, have close interrelationships with other critical

infrastructure, and can potentially impact communities and associated industries after the

disaster [Oh, Deshmukh and Hastak, 2010].

22
For Case 2 (community-driven off-site relocation), the provision of construction skills training to

households can positively affect the level of satisfaction. However, only 15% of the respondents were

able to join the training, as shown in Figure 6. Nonetheless, beyond the 500-hour sweat equity

agreement with the funding NGO, some men are also able to work as laborers or skilled workers in

the construction of the relocation site as temporary income source (more details provided in Section

7.2.2 below). Despite their limited coverage, the construction skills training and construction work

opportunities complemented the livelihood self-sufficiency variable, which in turn increased the level

of satisfaction of some respondents. In this sense, researchers studying the recovery of coastal

communities affected by the 2004 Asian tsunami such as Pomeroy et al. [2006] have also similarly

highlighted the importance of revitalizing livelihood options as a means of addressing the root cause

of vulnerability and thus building resilience in communities.

Meanwhile, for Case 3 (contractor-driven off-site relocation), no independent variable from the

collected survey data was found to be significant. As shown earlier in Figure 6, most respondents

from Case 3 did not receive any livelihood training or construction skills training assistance, although

all of them were selected for the relocation program and were able to receive relief goods. Further

investigation is warranted to better understand which type of assistance would address the needs of

contractor-driven off-site relocation case type.

Past relocation experiences in the Philippines indicate that greater participation of beneficiaries in

building and even designing their own houses have the following advantages [Ballesteros and Egana,

2012]: (1) beneficiaries can have freedom to address their need for bigger space as their family size

increases and (2) greater involvement and investment in housing construction increases community

participation and provides incentives for relocated beneficiaries to establish more effective

homeowners association. However, partially because most households have only recently moved in to

Case 2 (community-driven off-site relocation) and Case 3 (contractor-driven off-site relocation) sites,

the impact of using a community or contractor driven approach to permanent housing construction on

the level of satisfaction is not yet clear. How such factors will affect the case of Tacloban will be the

focus of future studies by the authors.


23
7.2 Social Challenges to Relocation

In both community- and contractor-driven relocation projects a number of critical social challenges

have already emerged at the time of writing, raising concerns about the long-term viability of the

projects. The following sections will discuss several other essential and challenging aspects that are

critical for a successful relocation, based on information from key-informants and discussions with

respondents after they had completed the questionnaires.

7.2.1 Accessibility and Basic Services

Permanent relocation sites are located more than 20 km north of downtown Tacloban, and takes about

an hour ride from the city center. As the sites are currently poorly served by public transport, residents

are faced with various related issues such as the cost of transportation, availability of livelihood and

accessibility of schools.

Each permanent house in Cases 2 (community-driven off-site relocation) and 3 (contractor-driven off-

site relocation) is to be provided with individual water and electricity connections. However, as the

construction of the relocation sites is still currently ongoing, these connections have not yet been

prepared. To address this gap, each household was provided with solar lamps by UNHCR in

partnership with IOM. In terms of water, residents of Case 2 currently have to buy water from an

outside source for PhP30/day or US$0.68/day (at least 10% of the daily incomes of those who have

work) or fetch it from a deep well around the site. It is important to note that many of the resettlement

beneficiaries previously lived as informal settlers in coastal barangays and are beneficiaries of

DSWDs 4P conditional cash transfer program which is mainly for families living below the countrys

poverty threshold (Case 2: 60% and Case 3: 33% of samples). As of the first semester of 2014, this

threshold was pegged at an average of PhP8,788 or US$198.6 per month for the basic food and non-

24
food needs of a family of 5 (Philippine Statistics Authority National Statistical Coordination Body,

2014).

According to the community leader of Case 2 (community-driven off-site relocation), in the long-term

residents will have to utilize their own money to obtain an electricity and water connection, which

cost PhP870 (US$20) and PhP2,500 (US$57), respectively. Water billing will be pegged at PhP250

per month (US$5.65/month), assuming that each household only consumes 10 cubic meters of water

or less. However, as of March 2015, water pipe connections are only available to some blockhouses

close to the road. Still, given the lack of income due to poor livelihood opportunities in the relocation

sites, acquiring electricity and water connections (when they are finally available) may remain a great

challenge to relocated households. The beneficiaries financial responsibilities after housing

completion and transfer are outlined in Table 5.

7.2.2 Livelihood Opportunities

The issue of livelihood is one of the most pressing for households in Cases 2 (community-driven off-

site relocation) and 3 (contractor-driven off-site relocation). As the sites are located far from

downtown Tacloban, access to jobs and other income-generating activities have become difficult for

residents. This was especially the case for women who used to work as market vendors and domestic

service providers, as there are no markets or business establishments near the relocation sites. On the

other hand, especially for Case 2 (community-driven off-site relocation), some men are able to work

in the construction of the permanent relocation site for PhP250/day (US$5.65/day) for laborers, and

PhP350-PhP400/day (US$7-9/day) for skilled workers.

While still staying in their transitional shelters, Case 2 residents were able to participate in a month-

long carpentry training provided by the International Labor Organization (ILO) upon the request of

the City Housing Office. ILO also provided participants with complete personal protective gears, tools

25
and a daily salary of PhP300 (US$6.8). The training was attended by both men and women (50%-50%

ILO requirement). However, a female resident of Case 2 admitted that:

Sumali lang ako para maka-ano yung mga eskwela ko, para makabili ng books... Hindi

na [ako nagtrabaho pagkatapos ng training]. Ang hirap, hindi ko kaya. Napasubo lang

ako dahil sa kahirapan. Para mapunuan ang pagkukulang, PhP1,800 yun per week. (I

only joined the training so my kids can continue studying, and buy their books. I did not

actually work after that because it is very difficult, I cannot take it. I just joined out of

poverty, so that we can address our financial needs. That was also PhP1,800 per week.).

Site observation by the authors also indicated that construction workers in the relocation site were

mostly men.

Thus, while women were able to participate in the training, they were unlikely to proceed to actual

construction work, leaving them with lesser livelihood opportunities compared to men. Still, as

construction will eventually be completed, longer-term livelihood opportunities need to be made

available to both sexes in general.

One of the main requirements for proper relocation is for adequate livelihood opportunities to be

available [Cernea, 1997]. However, as of March 2015, limited livelihood programs have been

implemented in Cases 2 (community-driven off-site relocation) and 3 (contractor-driven off-site

relocation). Most residents are still awaiting livelihood assistance from the Department of Social

Welfare and Development (DSWD) in partnership with the International Emergency and

Development Aid (IEDA) Relief Philippines. Under the DSWD scheme, residents will be provided

with livelihood capital based on the business proposals they have previously submitted. Based on the

beneficiaries own capacity and preference, they can propose to start or restart businesses in dress-

making, cooking and market-vending, among others. Funding will also depend on the proposal.

However, the program has already been delayed several times and frustration has grown among target

beneficiaries. Nonetheless, the Memorandum of Agreement between DSWD and the implementing

26
organization, International Emergency and Development Aid (IEDA), has already been signed. IEDA

is looking to receive funding and start program implementation in April 2015.

7.2.3 Educational Institutions

Although learning facilities like day care centers and even elementary schools are included in the

plans for relocation sites, construction of such infrastructure has not yet been started as of March

2015. In this sense, it appears that the priority is to complete the row houses and move in all

beneficiary families first, before commencing the construction of other facilities. For Case 2

(community-driven off-site relocation), an elementary school is located about 3 kilometers from the

site, which children can access by walking for an hour each way or by commuting for PhP10

(US$0.23) each way. However, for Case 3 (contractor-driven off-site relocation), participants of focus

group discussion revealed that their children have stopped going to school due to the inability to pay

for commuting fares and school allowance. At the same time, because their children do not go to

school anymore they have stopped receiving assistance from DSWD's 4P program. Based on

Administrative Order No. 16 of DSWD (2008), the 4P program grants households PhP300 or US$7

(for elementary students) or PhP500 or US$11 (for high school students) each month on the condition

that the children regularly attend classes.

Following all these challenges, many target beneficiaries who are still awaiting relocation might be

deterred from moving into their new houses after they have finally been completed. Similarly, they

might also accept the new houses but, at the same time, keep another house in informal settlements

close to downtown Tacloban (which might eventually result in a partial failure of the relocation

program, especially if informal settlements are allowed to be used by descendants of current

inhabitants, clearly a problem in an area with high fertility rates). In fact, as in Case 1a (on-site

reconstruction, inside the NDZ), many of the residents of these informal settlements had already

rebuilt their houses in their original locations, close to the coastline.

27
28
Conclusion

The post-Haiyan recovery effort in Tacloban has emphasized the need to incorporate the build-back-

better resilience-thinking to the housing reconstruction program in the affected coastal communities.

Under this framework, spatial solutions, which involved land use planning and building adaptation as

part of the multi-layer safety approach [Hoss et al. 2011], have been integrated with the shelter option

programs of government and non-government organizations through the no dwelling zone (NDZ)

policy. However, based on the survey conducted, some households were still living inside the NDZ.

This endangers the lives of these households against future disasters by failing to reduce the

vulnerability of communities in high-risk zones. In this regard, further studies on their choice of

shelter options (ex. why some households still decided to rebuild inside the NDZ) needs to be

investigated in detail.

The study also compared the challenges faced by three post-disaster housing reconstruction

approaches: Case 1 (on-site reconstruction), Case 2 (community-driven off-site relocation) and Case

3(contractor-driven off-site relocation). In terms of assistance options, Case 1 faced challenges

relating to lack of crucial assistance types (e.g. livelihood training) and, at the same time,

misallocation of others (e.g. shelter repair kit) as well as poor information dissemination.

Furthermore, this study found that prolonged disruption of water utility service and high severity of

damage to housing structures hampered the speed of household recovery in the affected areas.

On the other hand, for Case 2, insufficient livelihood training programs and opportunities posed a

challenge to the recovery of the displaced community, potentially affecting their perceived level of

self-sufficiency after the disaster. These factors were determined to be significant contributors towards

achieving sustainable socio-economic development for the communities in the relocation site.

In addition, for both Cases 2 and 3, issues such as delay in project implementation, problem in land

acquisition, and poor quality of construction work were apparent. These problems need to be

addressed for permanent relocation projects to bear positive impacts to the displaced household.

29
Finally, it is clear that there is significant room for improvements in the permanent housing programs

discussed, including setting up sound evaluation criteria for site acceptability prior to land acquisition

(e.g. accessible and safe), providing assistance to beneficiaries regarding sustainable livelihoods,

repairing critical infrastructure (e.g. transportation, electricity, water), pre-planning land acquisition to

expedite rebuilding process, improving the transparency and accountability of projects by keeping

target beneficiaries informed of the progress of reconstruction, and increasing the involvement of

beneficiaries. Further research into the lessons that can be learned from this disaster should be

undertaken in order to improve future recovery processes following natural disasters.

30
Acknowledgements

Funds for the field survey were mainly provided by the Graduate Program in Sustainability Science

Global Leadership Initiative (GPSS-GLI, The University of Tokyo), the Strategic Research

Foundation Grant-aided Project for Private Universities from Ministry of Education (Waseda

University), and a research grant of Tokyo Institute of Technology. The authors would also like to

acknowledge the contribution of International Emergency and Development Aid (IEDA) Relief

Philippines for their assistance to the survey.

31
References

Adger, W., Hughes, T.P., Folke, C., Carpenter, S.R., Rockstrom, J. [2005] Social-ecological

Resilience to Coastal Disasters, Science. 309 (5737), pp. 1035-1039.

AFP. [2011] At least 18 die in Philippines Typhoon, Herald Sun. Retrieved from:

http://www.heraldsun.com.au/archive/news/at-least-18-die-in-philippines-typhoon/story-

e6frf7lf-1226149286198. Accessed: 29 March 2015.

Asgary, A., Badri, A., Rafieian, M. and Hajinejad, A. [2006] Lost and Used Post-disaster

Development Opportunities in Bam Earthquake and the Role of Stakeholders, Proceedings

of the International Conference and Student Competition on Post-disaster Reconstruction:

Meeting Stakeholder Interests, Italy.

Ballesteros, M. and Jasmine Egana, J. [2012] Efficiency and Effectiveness Review of the

National Housing Authority [NHA] Resettlement Program, Philippine Institute for

Development Studies and Department of Budget. Accessed: 01 November 2014.

Barenstein, J.D. [2006] Housing Reconstruction in Post-earthquake Gujarat, HPN Network

Paper No. 54, Overseas Development Institute, London.

Berke, P., Kartez, J., Wenger, D. [1993] Recovery after Disaster: Achieving Sustainable

Development, Mitigation, and Equity, Disasters 17: pp. 93-109.

Bouraoui, D. and Lizarralde, G. [2013]. Centralized Decision-making, Users Participation

and Satisfaction in Post-disaster Reconstruction: The Case of Tunisia, International Journal

of Disaster Resilience in the Built Environment, Vol. 4, No. 2, pp. 145-167.

Brand, S., Blelloch, J. W. [1973] Changes in the Characteristics of Typhoons Crossing the

Philippines, J. Appl. Meteor., 12, 104109.

Bricker, J. and Roeber, V. [2015] Mechanisms of Damage During Typhoon Haiyan: Storm

Surge, Waves and Tsunami-Like Surf Beat, Proc. Of 36th IAHR World Congress, 28 June-

3 July, The Hague, Netherlands.

32
Brown, S. [2013, November] The Philippines is the Most Storm-exposed Country on Earth,

TIME Magazine. Retrieved from: http://world.time.com/2013/11/11/the-philippines-is-the-

most-storm-exposed-country-on-earth/. Accessed 29 March 2015.

City Government of Tacloban [2014] The Tacloban Recovery and Rehabilitation Plan:

Shelter, Republic of the Philippines, UN-HABITAT and UNDP.

Cohen, J.W. [1988] Statistical Power Analysis for the Behavioral Sciences 2nd edition,

Hillsdale, NJ: Lawrence Erlbaum Associates.

Coppola, D.P. [2007] Introduction to International Disaster Management, Elsevier, Boston.

Cutter, S., Barnes, L., Berry, M., Burton, C., Evans, E., Tate, E. and Webb, J. [2008] A Place-

based Model for Understanding Community Resilience to Natural Disasters, Global

Environmental Change. Vol. 18, Issue 4, pp. 598-606.

Department of Environment and Natural Resources Region VIII [2014] DENR Delineates 20

and 40 meters easement of coast lines affected by Typhoon Yolanda, Department of

Environment and Natural Resources Region VIII, Palo, Leyte, Philippines.

Department of Environment and Natural Resources (DENR), Department of Interior and

Local Government (DILG), Department of National Defense (DND), Department of Public

Works and Highways (DPWH) and Department of Science and Technology (DOST) [2014]

Adoption of Hazard Zone Classification in Areas Affected by Typhoon Yolanda (Haiyan)

and Providing Guidelines for Activities Therein, Memorandum Circular 2014-01.

Department of Science and Technology (DOST) [2014, May] DOST, DENR Launch Info

Center for Yolanda Rehab, Retrieved from: http://www.dost.gov.ph/index.php/knowledge-

resources/news/34-2014-news/85-dost-denr-launch-info-center-for-yolanda-rehab. Accessed:

29 March 2015.

Department of Social Welfare and Development (DSWD) [2008] Guidelines on the

Implementation of Pantawid Pamilyang Pilipino Program(4Ps), Retrieved from:

http://www.fo1.dswd.gov.ph/wp-content/uploads/2013/07/AO-No-16-s-2008.pdf. Accessed

29 March 2015.
33
DSWD, International Organization of Migrants (IOM), Internal Displacement Monitoring

Centre (IDMC) and SAS [2014] The Evolving Picture of Displacement in the Wake of

Typhoon Haiyan: An Evidence-based Overview, Retrieved from:

http://www.iom.int/files/live/sites/iom/files/Country/docs/The-Evolving-Picture-of-

Displacement-in-the-Wake-of-Typhoon-Haiyan.pdf. Accessed 29 March 2015.

Del Castillo, Butch [2013] Wind Zones and the Building Code, Business Mirror, Manila,

Philippines. Retried from: http://businessmirror.com.ph/index.php/en/news/opinion/23014-

wind-zones-and-the-building-code. Accessed 17 January 2014.

Environmental Science for Social Change (ESSC) [2014]. Rapid Assessment of the

Performance of Post-Disaster Housing Reconstruction Approaches, Retrieved from:

http://essc.org.ph/content/wp-content/uploads/2014/10/ESSC-final-report_30June-2014.pdf.

Accessed 29 March 2015.

Esteban, M., Stromberg, P. Gasparatos, A. and Thomsom-Pomeroy, D. [2012] Global

Warming and Tropical Cyclone Damage in the Philippines, Journal of Climate Research,

Vol. 56: 5160, 2013.

Esteban, M. Tsimopoulou, V., Mikami, T., Yun, N. Y., Suppasri, A. and Shibayama,T. [2013]

Recent Tsunami Events and Preparedness: Development of Tsunami Awareness in Indonesia,

Chile and Japan, International Journal of Disaster Risk Reduction, Vol. 5, pp. 84-97.

Esteban, M., Nguyen D. T., Takagi, H., Valenzuela, P. and Tam, T.T. [2014] Storm Surge and

Tsunami Awareness and Preparedness in Central Vietnam, In Coastal Disasters and Climate

Change in Vietnam: Engineering and Planning Perspectives, Nguyen D. T., Takagi, H. and

Esteban, M. editors, Elsevier.

Esteban, M., Takagi, H. and Shibayama, T., [2015a] Coastal Disasters: Lessons Learnt for

Engineers and Planners, Elsevier (in print)

Esteban, M., Thao, N. D., Takagi, H., Tsimopoulou, V., Mikami, T., Yun, N.Y. and Suppasri,

A. [2015b] The Emergence of Global Tsunami Awareness: Analysis of Disaster Preparedness

34
in Chile, Indonesia, Japan and Vietnam in Coastal Disasters: Lessons Learnt for Engineers

and Planners. Esteban, M., Takagi, H. and Shibayama, T. (eds.). Elsevier.

Esteban, M., Onuki, M., Ikeda, I and Akiyama, T. [2015c] Reconstruction Following the

2011 Tohoku Earthquake Tsunami: Case Study of Otsuchi Town in Iwate Prefecture, Japan

in Coastal Disasters: Lessons Learnt for Engineers and Planners. Esteban, M., Takagi, H.

and Shibayama, T. (eds.). Elsevier.

Esteban, M., Valenzuela, V. V. Namyi, Y., Mikami, T., Shibayama, T., Matsumaru, R., Takagi,

H. Thao, ND., de Leon M., , Oyama, T. Nakamura, R. [2015d] Typhoon Haiyan 2013

Evacuation Preparations and Awareness, J-SustaiN [accepted]

Folke, C. [2006] Resilience: The Emergence of a Perpective for Social-ecological Systems

Analyses, Global Environmental Change 16 (3), 253-267.

Harrington, L.S.B. [2005] Vulnerability and Sustainability Concerns for the US High

Plains, In Rural Change and Sustainability: Agriculture, the Environment and Communities.

S.J. Essex, A.W. Gilg, and R. Yarwood (eds.). Cambridge, MA: CABI Publishing, pp. 169-

184.

Karunasena, G. and Rameezdeen, R. [2010] Post-disaster Housing Reconstruction:

Comparative Study of Donor vs Owner-driven Approaches, International Journal of

Disaster Resilience in the Built Environment, Vol. 1, No. 2, pp. 173-191.

Kates, R.W., Colten, C.E., Laska, S. and Leatherman, S.P. [2006] Reconstruction of New

Orleans after Hurricane Katrina: A Research Perspective, Proceedings of the National

Academy of Sciences of the United States of America, 103, 14653-14660.

Klein, R.J.T., Nicholls, R.J., Thomalia, F. [2003] Resilience to Natural Hazards: How Useful

is this Concept? Environmental Hazards, 5 (1-2), 35-45.

Kubota, H., and J. C. L. Chan [2009] Interdecadal Variability of Tropical Cyclone Landfall

in the Philippines from 1902 to 2005, Geophys. Res. Lett., 36, L12802,

doi:10.1029/2009GL038108.

35
Lagmay, A. M. F., Agaton, R. P., Bahala, M. A. C., Briones, J. B. L. T., Cabacaba, K. M. C.,

Caro, C. V. C., Dasallas, L. L., Gonzalo, L. A. L., Ladiero, C. N., Lapidez, J. P., Mungcal, M.

T. F. Puno, J. V. R., Ramos, M. M. A. C., Santiago, J., Suarez, J. K. and Tablazon, J. [2015]

Devastating storm surges of Typhoon Haiyan, Intl. Journal of Disaster Risk Reduction, 11,

pp. 1-12.

Lapidez, J.P., Suarez, J. K., Tablazon, J., Dasallas, L., Gonzalo, L.A., Santiago, J., Cabacaba,

K.M., Ramos, M.M.A., Lagmay, A.M.F., Malano, V. [2014] Identification of Storm Surge

Vulnerable Areas in the Philippines through Simulations of Typhoon Haiyan-Induced Storm

Surge Using Tracks of Historical Typhoons, DOST-Project NOAH Open-File Reports Vol. 3

(2014), pp. 112-131, ISSN 2362 7409.

Leelawat, N, Mateo, C. M. R., Gaspay, S. M., Suppasri, A., Imamura, F. [2013] Filipinos

Views on the Disaster Information for the 2013 Super Typhoon Haiyan in the Philippines,

International Journal of Sustainable Future for Human Security, J-SustaiN. Vol. 2 No. 2 pp.

61-73.

Lizarralde, G., Johnson, C. and Davidson, C.H. [2009] Rebuilding After Disasters: From

Emergency to Sustainability, Taylor & Francis, London.

Matsumaru, R. [2015] Reconstruction from the Indian Ocean Tsunami Disaster: Case study

of Indonesia and Sri Lanka and the philosophy of Build Back Better, in Coastal Disasters:

Lessons Learnt for Engineers and Planners. Esteban, M., Takagi, H. and Shibayama, T.

(eds.). Elsevier.

Mileti, D.S. [1999] Disaster by Design: A Reassessment of Natural Hazards in the United

States, Joseph Henry Press, Washington, D.C.

Mori, N., Kato M, Kim S., Mase H., Shibutani Y., Takemi T., Tsuboki K., Yasuda T.

[2014] Local Amplification of Storm Surge by Super Typhoon Haiyan in Leyte Gulf,

Geophys. Res. Lett., 41, 51065113, doi:10.1002/2014GL060689.

36
National Water Plan of the Netherlands, Retrieved from:

http://english.verkeerenwaterstaat.nl/english/Images/NWP%20english_tcm249-274704.pdf.

Accessed 10 August 2012.

National Disaster Risk Reduction and Management Council (NDRRMC) [2014] Effects of

Typhoon YOLANDA (HAIYAN), Situation Report No. 104, Retrieved from:

http://www.ndrrmc.gov.ph/attachments/article/1329/Effects_of_Typhoon_YOLANDA_(HAI

YAN)_SitRep_No_104_29JAN2014_0600H.pdf. Accessed 19 Jan 2014.

Oh, E.H., Deshmukh, A. and Hastak, M. [2010] Disaster Impact Analysis based on Inter-

relationship of Critical Infrastructure and Associated Industries, International Journal of

Disaster Resilience in the Built Environment, Vol.1, No. 1, pp. 25-49.

Philippine National Government- Department of Environment and Natural Resources [1976]

Presidential Decree 1067: Water Code of the Philippines, Chapter 4, Article 51. Retrieved

from: http://www.lawphil.net/statutes/presdecs/pd1976/pd_1067_1976.html. Accessed 29

March 2015.

Philippines Shelter Cluster [2014a]. Inter-Cluster Advisory to the HCT on the Provision of

Assistance in Proposed No Dwelling Zones, Retrieved from:

https://www.sheltercluster.org/sites/default/files/docs/Advisory%20Note%20-%20No

%20Dwelling%20Zones.pdf. Accessed 29 March 2015.

Philippines Shelter Cluster [2014b] Shelter Sector Response Monitoring: Final Report

Monitoring Assessment 2, Retrieved from

https://www.sheltercluster.org/sites/default/files/docs/reach_phl_report_haiyan_sheltersectorr

esponsemonitoring2_sep2014_0.pdf. Accessed 29 March 2015.

Presidential Assistant for Rehabilitation and Recovery (PARR) [2014] No Build Zone Policy

Not Recommended in Yolanda-affected areas, Retrieved from:

http://www.gov.ph/2014/03/14/parr-no-build-zone-policy-not-recommended-in-yolanda-

affected-areas/. Accessed 29 March 2015.

37
Quarantelli, E.L. [1982] General and Particular Observations on Sheltering and Housing in

American Disasters, Disasters 6: 277-281.

Palliyaguru, R. and Amaratunga, D. [2011] Linking Reconstruction to Sustainable Socio-

economic Development, in Post-disaster Reconstruction of the Built Environment:

Rebuilding for Resilience. Wiley-Blackwell, Chapter 15.

Paul, B.K. [2011] Disaster Cycles: Response and Recovery, In Environmental Hazards and

Disasters: Contexts, Perspectives and Management. Wiley- Blackwell. Chapter 6.

Philippine Statistics Authority National Statistical Coordination Body [2014] Annual

Poverty Indicators Survey, Republic of the Philippines.

http://www.nscb.gov.ph/pressreleases/2015/PSA-20150306-SS2-01_poverty.asp, Accessed

23rd, April, 2015

Pomeroy, R., Ratner, B., Hall, S., Pimoljinda, J. and Vivekanandan, V. [2006] Coping with

Disaster: Rehabilitating Coastal Livelihoods and Communities, Marine Policy, Vol. 30, Issue

6, pp. 766-793.

Project NOAH [2013] Philippine Storm Surge History, Retrieved from:

http://blog.noah.dost.gov.ph/2013/11/22/philippine-storm-surge-history/. Accessed 29 March

2015.

Rappler. [2014] NOAH: More Areas Warned versus Ruby Storm Surges, Retrieved from:

http://www.rappler.com/move-ph/issues/disasters/typhoon-ruby/77149-20141206-5am-ruby-storm-surge-warnings . Accessed

29 March 2015.

Ranada, P. [2014] After Yolanda: Are We More Prepared for Storm Surge? Retrieved from:

http://www.rappler.com/nation/74089-yolanda-year-after-storm-surge-advisories. Accessed 29 March 2015.

Regala, M. [2014] Standards for Housing Design and Reconstruction: Relevance or

Applicability in View of New Normal, National Housing Authority, Republic of the

Philippines. Retrieved from http://essc.org.ph/content/wp-

content/uploads/2014/04/Day2Session3A_REGALA_NHA_Standards-for-Housing-Design-

Construction.pdf. Accessed 29 March 2015.

38
Republic Act (RA) 10121. [2010]. Republic Act 10121: Philippine Disaster Risk Reduction

and Management Act of 2010, Manila, Philippines: Republic of the Philippines.

Schiermeier, Q. [2013] Did Climate Change cause Typhoon Haiyan, Nature.

doi:10.1038/nature.2013.14139.

Schilderman, T. [2010] Putting People at the Center of Reconstruction. From: Building Back

Better: Delivering People-centered Housing Reconstruction at Scale, Lyons, M.,

Schilderman, T. and Boano, C. (eds.). Practical Action Publishing, UK, p. 30.

Shibayama, T., Esteban, M., Nistor, I., Takagi, H., Danh Thao, N., Matsumaru, R., Mikami,

T., Aranguiz, R., Jayaratne, R. and Ohira, K. [2013] Classification of Tsunami and

Evacuation Areas, Journal of Natural Hazards, 67 (2), 365-386

Stromberg, P., Esteban, M. and Gasparatos, A. [2011] Climate Change Effects on Mitigation

Measures: the Case of Extreme Wind Events and Philippines biofuel plan, Journal of

Environmental Science & Policy, Vol. 14, 8, pp 1079-1090.

Shibayama T., Matsumaru R., Takagi H., Leon M., Esteban M., Mikami T., Oyama T.,

Nakamura R. [2014] Field survey and analysis of storm surge caused by the 2013 Typhoon

Yolanda (Haiyan), Journal of Japan Society of Civil Engineers, Ser. B3 (Ocean

Engineering), Vol. 70 [2014] No. 2, pp.1212-1217,

http://dx.doi.org/10.2208/jscejoe.70.I_1212

Smith, G. and Wenger, D. [2006] Sustainable Disaster Recovery: Operationalizing an

Existing Framework, In Handbook of Disaster Research. H. Rodriguez, E. Quarantelli, & R.

Dynes (eds.). Springer, New York, pp. 234257.

Suzuki, S. [2012] Tsunami resilient community development from discussions with affected

people Advances in Coastal Disasters Risk Management, Lessons from the March 2011

Tsunami and preparedness to the climate change impact Seminar, Sendai, Japan, 7 th-8th June

2012.

Takagi, H., De Leon, M., Esteban, M., Mikami, T., and Nakamura, R. [2015] Storm Surge

due to 2013 Typhoon Yolanda (Haiyan) in Leyte Gulf, the Philippines in Coastal Disasters:

39
Lessons Learnt for Engineers and Planners. Esteban, M., Takagi, H. and Shibayama, T.

(eds.). Elsevier.

Takagi, H., Esteban, M., Shibayama, T., Mikami, T., Matsumaru, R., de Leon M.,

Thao, ND., Oyama, T. [2014], Track Analysis, Simulation and Field Survey of the

2013 Typhoon Haiyan Storm Surge, Journal of Flood Risk Management,

DOI:10/1111/jfr3/12136

Takagi, H., Esteban, M., Shibayama, T., Mikami, T., Matsumaru, R., de Leon M., Thao, ND.,

Oyama, T., [2015] Track Analysis, Simulation and Field Survey of the 2013 Typhoon Haiyan

Storm Surge, Journal of Flood Risk Management DOI:10/1111/jfr3/12136

Takagi, H., Kashihara, H., Esteban, M. and Shibayama, T. [2011] Assessment of Future

Stability of Breakwaters under Climate Change, Coastal Engineering Journal (CEJ), Vol.

53, No. 1, pp. 21-39

Tajima Y., Yasuda T., Pacheco B., Cruz E., Kawasaki K., Nobuoka H., Miyamato M. Asano

Y., Arikawa T., Origas N. M., Aquino R., Mata W., Valdez J., Briones F. [2014] Initial report

of JSCE-PICE Joint Survey on the Storm Surge Disaster caused by Typhoon Haiyan,

Coastal Eng. J., Vol. 56, No.1, DOI: 10.1142/S0578563414500065

Thiruppugazh, V. [2007] Post-disaster Reconstruction and the Window of Opportunity: A

Review of Select Concepts, Models and Research Studies, JTCDM Working Paper Series

No. 3, Tata Center for Disaster Management, India.

Tobin, G.A. and Montz, B.E. [1997] Natural Hazards: Explanation and Integration, New

York: The Guilford Press.

Tsimopoulou, V., Vrijling, J.K., Kok, M., Jonkman, S.N., Stijnen, J.W. [2013] Economic

Implications of Multi-layer Safety projects for Flood Protection, Proc. ESREL Conference,

Amsterdam.

Tsimopoulou, V., Jonkman S.N., Kolen, B., Maaskant, B., Mori, N., Yasuda, T. [2012] A

Multi-layer Safety Perspective on the Tsunami Disaster in Tohoku, Japan, Proc. Flood Risk

2012 conference, Rotterdam.

40
United Nations Development Program (UNDP) [2006]. Post-conflict Reconstruction of

Communities and Socio-economic Development, New York: Bureau of Crisis Prevention

and Recovery.

Wisner, B. et al. [2004] At Risk: Natural Hazards, Peoples Vulnerability and Disasters,

Routledge, London.

World Bank [2014]. Recovery and Reconstruction Plan in the Aftermath of Typhoon Haiyan

(Yolanda): Summary of Knowledge Briefs. Washington DC, USA.

41

Potrebbero piacerti anche