skip to main content
10.1145/3613904.3642581acmconferencesArticle/Chapter ViewFull TextPublication PageschiConference Proceedingsconference-collections
research-article
Open Access

Towards More Sustainable Interactive Textiles: A Literature Review on The Use of Biomaterials for eTextiles.

Published:11 May 2024Publication History

Abstract

The development of functional fibres, active materials, and flexible electrical components has introduced new ways of embedding interactive capabilities within textiles. However, this seamless integration poses challenges in terms of materials, disassembly, and disposal, revealing an urgent need to address the issue of sustainability when creating new electronic textiles. Authors have proposed eco-design guidelines that emphasise the use of renewable and biodegradable materials. Despite these recommendations, the potential of biomaterials in eTextiles remains largely unexplored. This integrative literature review showcases how biomaterials emerged as catalysts for expanding possibilities within eTextiles and HCI, not only through their environmental sustainability but also through their dynamic and transformative nature, fostering a realm of novel interactive experiences. We suggest the potential of developing fully bio-based eTextile systems, the need for broader sustainability and aesthetic studies, the relevance of DIY methods, and the urgency of textile knowledge integration.

Skip 1INTRODUCTION Section

1 INTRODUCTION

Textiles with integrated electrical capabilities, i.e. electronic textiles (or eTextiles), are attracting increasing research interest as they represent a flexible, conformable platform for wearable and interactive applications. The potential application areas cover a broad spectrum of various use contexts, ranging from stroke rehabilitation [158] and tracking physical training performance [133] to smart home interfaces [60]. Due to their versatile application areas, and how in combination with technological advancements in material science and flexible electronics they enable more comfortable electronics integration, eTextile products are also gaining increasing consumer interest. This has led to notable growth in the eTextile market, which is expected to reach US 1.4 billion by 2030 [59]. As the eTextile market sits at the intersection of the textile and electronics sectors, the market infiltrates and integrates aspects of both. Although it introduces an intriguing interdisciplinary playground for developing sophisticated wearable devices, this convergence also entails significant waste management challenges that both fields are facing [7, 112]. More specifically, the integration of electrical components and textile materials introduces new additional challenges related to material resources and waste management [144], which are yet to be solved. Hence, the sustainability concerns stemming from eTextile manufacturing, utilisation, and disposal urgently need addressing, which calls for integrating biodesign guidelines and aiming for more sustainable options, such as prioritising renewable and recycled materials [76].

Figure 1:

Figure 1: Biomaterials and their uses for the creation of different eTextile system components

In practice, eTextile development involves a high degree of electrical component integration into textiles to increase the comfort, flexibility, robustness, reliability and maintenance of eTextile applications. This quest for seamless integration, and subsequently for better user acceptance [120], has led to a wide range of textile production methods. Additive textile fabrication methods such as lamination [145] have been used to attach circuits to ready-made textile substrates. Lamination typically uses thermoplastic polyurethane (TPU) film for bonding. Although it enables the fabrication of stretchable and reliable eTextile systems [145] and the disassembly of electronics for recycling, the approach includes several additional production phases, and uses more material resources. Electrically active yarns and fibres [87], flexible electronics [27], and yarns with embedded components [56, 57] can also be integrated into textile structures during the textile substrate construction process, making them inseparable parts of the fabric. For instance, using silver-coated nylon yarns has led to the development of intricate multilayer woven sensors and circuitry [114], offering possibilities to fully embed eTextile systems into woven textile structures. Sealing the electronics inside the textile protects the electrical system from external exposure, as well as the wearer from the system. This approach also allows the construction of eTextile systems that do not compromise the fabrics’ visual and tactile aesthetic qualities, which is essential for product adoption [72]. However, such an approach introduces several sustainability issues. First, such eTextiles combine raw textile materials and electronics crafted from scarce resources, heavy metals, and toxic chemicals [44]. Second, as the electrical properties are embedded in the woven structure, the conductive yarns cannot be disassembled from the fabric, hampering its recyclability. Additionally, system-level integration may also involve the utilisation of potentially hazardous substances, such as flexible energy storage devices [124].

Augmenting textiles with electrical capabilities also has a significant impact on products’ life cycles. Once in use, eTextile products are likely to be situated as mass consumer goods in the fashion and electronics sectors, promoting evolving styles and technological obsolescence [44], as well as cost-effective applications that are discarded rather than repaired. In addition, complex constructions, a combination of fabrication techniques such as weaving and soldering, and uncommon maintenance procedures reduce the repairability of such products. At the end-of-life stage, the intricate integration and diverse waste streams pose challenges to recyclability. As the eTextile market is still evolving, the industry suffers from a lack of standardisation, product traceability, and limited process feasibility. Currently, the WEEE Directive 2012/19/EU [28] regulates eWaste in Europe, while the OEKO-TEXr Standard 100 certification [131] ensures the elimination of harmful substances from textile supply chains. However, no specialised regulations exist for eTextiles. This absence of regulations has limited the deployment of adequate recyclability systems, and consequently, the recycling rate has remained low in Europe, where only a third of entities engaged in electronics and textile collection and recycling have established specialised processing strategies for eTextiles [34]. Without processing strategies, removing and reusing electrical parts such as conductive yarns is unfeasible. As a result, the products are shredded into smaller pieces in the recycling stage without being separated into constituent components, and the scarce materials embedded in the eTextile systems cannot be salvaged [44].

Figure 2:

Figure 2: Different stages of life cycle thinking for circularity in eTextiles

Thus, eTextile sustainability faces manifold challenges, which revolve around issues related to the union of electronics and textile materials, as well as to the life cycle of eTextile products. Reciprocally, addressing these challenges requires a holistic approach in which integrating life cycle thinking and circular economy models is pertinent [93]. Several eTextile scholars have investigated these challenges by focusing on different aspects of products’ life cycles and introducing eco-design guidelines [76] (Figure 2).

To start with the design and development phase, more attention must be paid to the users’ needs. This could be achieved with a purpose-led electronics and textile system design process in which the textile design process is steered towards considering the meaningful use of the end applications [152]. Involving users actively in the design process by employing co-design methodologies engenders user affinity, thereby enhancing product longevity [108]. To this end, a multi-tiered approach comprising reflective/expressive, behavioural/functional, and visceral/aesthetic dimensions has been suggested, to shape the framework for crafting sustainable smart textiles and clothing [44]. During the manufacturing phase, more sustainable eTextile production could be approached through a systematic design framework that adopts 4R eTextile design principles involving repair, recycling, replacement, and reduction [125]. In the use phase, sustainable consumption and product longevity can be achieved by creating lease-service systems [145], enabling repairability via modular solutions [56], establishing compatibility standards [76], avoiding functional obsolesce by promoting repairability [144], and advancing consumer awareness [108]. On the later stage the end-of-life, proper labelling, modular design, and removable components can provide solutions for increasing the recyclability of eTextile products [76, 145, 156]. Additionally, extending the scope of the WEEE Directive to cover eTextiles would contribute to legal frameworks for efficient recyclability [34].

Going back to the initial stage of material acquisition phase, the authors emphasise seeking alternatives sourced from renewable, non-toxic, ideally local, and biodegradable materials [52, 76, 124]. This principle is exemplified in Van der Velde’s Life Cycle Assessment [142], in which material selection emerges as the most beneficial strategy for decreasing the environmental impact of smart textile products. Although rare, a few studies have explored the realm of green electronics for eTextiles and wearables, highlighting the utility of biodegradable metals, conductive polymers, organic semiconductors, graphene, starch, and gallium-based liquid metals [8, 32, 124, 158]. Nonetheless, sustainable material alternatives have not been extensively explored, despite their direct implications for manufacture, usage and disposal, and their potential to significantly mitigate the environmental impact.

To enable a better understanding of the options for sustainable eTextile production, this paper delves into a specific category of materials with substantial promise as substitutes for the conventional plastics and metals ubiquitous in eTextile production: biomaterials. This review seeks to identify what and how biomaterials have been utilised in the creation of eTextiles. It also draws attention to how biomaterials can contribute to making human–material interaction applicable in eTextile development within human–computer interaction (HCI) research, and the extent to which these materials can contribute to more sustainable eTextile production.

Skip 2BACKGROUND Section

2 BACKGROUND

2.1 Biomaterials

On the basis of the word’s prefix, biomaterials can be understood as materials connected to living entities. In the field of biomedical research, they refer to various synthetic and natural materials, such as polymers or ceramics, used in the treatment of any tissue, organ or function in the human body [134]. In the domain of design, the interest in biomaterials is pushed by sustainability-driven trends, such as the war on plastics or veganism, together with the aim to incorporate living organisms into products [84], resulting in the linkage of material science, biology, design, arts, and crafts. Within this context, biomaterials are mainly understood as bio-based materials with a certain percentage of biological components [84]. Depending on the literature source, the percentage required to fulfil this definition varies from 10% upwards to 100% [84, 129]. This divergence influences the level of biodegradability, which, according to Materiom [129], must be 100% guaranteed. Contributing to the ongoing definition process, Weiss et al. [2002] introduce a territorial dimension to biomaterials. They assert that these materials should only be from local biomass and ‘subscribe to the principles of circular economy and green chemistry, which nourish the territory in its decomposition process’ [150]. For the scope of this review, we consider biomaterials as materials that are 100% biobased and capable of complete biodegradation [130, 150]. They can originate from cultivating biological raw resources, such as mycelium, or from waste biomass, such as chitin.

The potential of biomaterials as better alternatives for product development stems from their role in future circular economies to help achieve some of the United Nations sustainable development goals [102]. According to Rosenboom et al. [2022], the introduction of bioplastics, one of the largest categories in biomaterials, could contribute to ‘diverting from fossil resources, introducing new recycling and degradation pathways, and using less toxic reagents and solvents in production processes’. Taking this into account, biomaterials and bioplastics are not, by default, more sustainable. Thus, life cycle assessments have to be conducted to determine the impact of every step, from feedstock harvesting to end-of-life [119].

In the case of textile and fashion industries, a first generation of companies have developed biomaterials such as mycelium leather-like sheets [100] or non-woven textile materials from pineapple leaf fibre [3]. Lee et al. [2020] have compiled a comprehensive list of specialised vocabulary and of these innovations [84]. In the electronics field, research has shown that biomaterials such as silk fibroin, cellulose and chitin can be used in flexible electronic devices in health monitors and biosensors [132].

Particularly in HCI, interest in biomaterials and integrating living organisms for eco-conscious biodigital interfaces is surging. This interest is framed, firstly, by a growing body of research on a material-driven approach that has contributed to a better understanding of how systems are crafted [38] and how the material properties of the things we interact with influence user experiences [69]. A second motivation stems from sustainability concerns within the HCI community. Researchers aim to incorporate recyclable and compostable materials into their prototyping processes, such as clay for 3D printing [19]. They are also conducting life cycle assessment studies in digital fabrication [83] and connecting HCI practice to sustainable development goals [54].

So far, materials such as mycelium [46, 80], food-waste bioclays [15], bacterial cellulose [106], iridescent bacteria [78], and bioluminescent algae [10, 107] have sparked innovation in interaction interfaces in HCI. Various methodological approaches have been employed to examine material properties, characteristics, agency, and their relationship with the human body. Crafting and DIY techniques have been embraced to explore biomaterials in electronics [80], emphasising care, patience, and comprehension during design stages [16], and to transfer educational knowledge [20]. Autoethnographic research has studied ‘life as shared experience’ and more-than-human design perspectives [78, 106]. These experiences have led the way to promoting organism-centred design and ethical considerations of the well-being of non-human organisms [107].

This review focuses on the search for and analysis of the available research on biomaterials applied in creating eTextile system elements by integrating studies framed in the field of eTextiles and soft wearable electronics.

2.2 Biomaterial Definitions

2.2.1 Renewable sources.

Compared with non-renewable petroleum-based materials and metals commonly used in eTextiles, biomaterials originate from renewable sources, meaning they can be part of circular production processes. They can be obtained through the cultivation of biological raw material or the processing of waste biomass. ‘Biorefineries’ convert renewable bio-based feedstock into useful chemicals, such as cellulose from straw or alginate from algae. In order to reduce their carbon footprint, their production should incorporate generative feedstocks derived from regenerative industries and processes [129], they should not compete with food production [119], and the manufacturing processes need to be energy efficient.

2.2.2 Biodegradability.

Biodegradation is the chemical transformation of a material carried out by microorganisms naturally found in the environment: during this process, these materials are converted into natural small substances such as water, carbon dioxide and biomass, without the need for artificial additives. Importantly, the biodegradability of a material is determined by its chemical composition rather than its source: ‘100 percent biobased plastics may be non-biodegradable, and 100 percent fossil based plastics can biodegrade.’ [18]. The rate and success of biodegradation are influenced by various factors, including environmental conditions (e.g. temperature and type of soil), the specific material in question and its intended use; this is why it is important to specify the difference to composting, a biodegradation process that takes place under controlled conditions in a given time frame and can be carried out at home, in a laboratory or in an industrial plant. Thus, the main difference is that biodegradation is a property of a material, whereas composting is a specific process [43].

2.2.3 Biocompatibility.

This is the ability of a material, or substance, to generate a biological response in a determined situation, without producing harmful effects. It also refers to the compatibility of the substance with the living environment [153]. In most cases, biomaterials can be innocuous and non-toxic for biological systems, particularly the human body. This characteristic makes them suitable for interactive textile products that need to be in close contact with the body, such as wearable devices or smart clothing [42].

2.2.4 Biopolymer.

Naturally occurring long-chain molecules e.g. polysaccharides, proteins, DNA [153].

2.2.5 Bioplastic.

Currently used in industry, bioplastics are defined as bio-based, biodegradable, or both [18].

This review will focus on the search and analysis of available research about biomaterials applied in creating eTextiles systems elements, integrating studies framed in the field of eTextiles and soft wearable electronics.

Skip 3METHOD Section

3 METHOD

This integrative literature review [151] addresses the lack of a comprehensive overview of how biomaterials have been used in more sustainable eTextile research and development. Due to the multidisciplinary nature of biomaterial research, we gathered the literature included in this review from multiple outlets and various disciplines. Furthermore, to gain a deeper understanding of the researched materials, their use, benefits and challenges, the sources were analysed and discussed from two perspectives: eTextile design and material science.

In the first phase, we searched leading journals and selected conference proceedings from the fields of textile design, HCI, electrical engineering, and material science (see Section 4.1 for more details). Our search keywords were ‘electronic textiles’ and ‘bio’, in the paper´s names or abstracts. The first keyword excluded smart textiles with no electrical capability, together with wearable technologies and green electronics that did not include textile substrates. The word ‘bio’ gave us flexibility to find materials that were later filtered by ‘bio-based’ and ‘biodegradable’. Materials that had only biodegradability traits, for example, were excluded. Due to the current extensive research on carbon and graphene fibres and textiles, an extra inclusion criterion was considered, particularly for carbon-based materials. Only papers with a clear statement of sustainability were included, for which we used the words ‘sustainable’, ‘environmentally friendly’, and ‘green’. We complemented the search by conducting citation tracking on the included papers, using the snowballing technique [154]. This resulted in 33 sources being included in the corpus.

For categorisation, we used a concept matrix to group and analyse the papers. The selected concepts corresponded to eTextile system components: power sources, conductive traces, substrates, microcontrollers, sensors, effectors, passive elements (resistors, capacitors, inductors) and cases (Figure 6). This enabled us to identify which kind of components biomaterials are mostly being used for and how they are integrated into these systems. Secondly, we analysed the materials on the basis of their origin, manufacturing methods, possible formats, combination with other materials, and biocompatibility. In addition, the different authors highlighted their interaction capabilities on the basis of the presented results. Finally, we searched for patterns of biodegradability, life cycle assessments, user studies and scale.

Skip 4RESULTS Section

4 RESULTS

4.1 General Summary of Results

Figure 3:

Figure 3: Distribution of articles over different publication outlets. In total, 19 results where found in the field of Material Science (dark red), 8 in HCI (red), 6 in Electrical Engineering (orange) and 2 in Textile related outlets (yellow).

Figure 4:

Figure 4: Material categories found with their corresponding amount of publications. Carbon-based materials are the most studied ones for the development of eTextiles.

The final search covered a total of 33 publications. These sources were found in 19 different outlets following the search literature review protocol. As illustrated in Figure 3, the majority of the research contributions emerged through material science outlets, followed by those related to HCI, electrical engineering and lastly, textiles. Particularly in textiles, we observed a clear gap. Even though research on smart textiles and electronic textiles was easily available, only two papers that match the inclusion criteria were found in a search of five leading textile journals and conference proceedings. Therefore, the following results are currently lacking more presence of knowledge originating from the textile field.

Figure 5:

Figure 5: Publications showing on the body examples, user studies and biodegradability tests.

A total of 9 different types of materials categories were found, with carbon-based materials being clearly predominant in the research outputs (Figure 4). This may be due to the extensive existing research on the capacity of graphene, carbon black (CB) and carbon nanotubes (CNT) to conduct electricity, a key feature of eTextiles. For the other eight materials we found only one to three publications, portraying a considerable research gap. During the search, conductive polymers such as PPy and PEDOT:PSS arose: their capacity to conduct electricity and biocompatibility has led to extensive studies for their use in electronic textiles and flexible electronics [138]. Nevertheless, they are not included in this review as they do not come from renewable sources nor can they biodegrade.

All the publications showed material development on a Technology Readiness Level 1 [101], portraying basic observed principles and formulating initial technology concepts and applications. We explain the fabrication processes and material details for all the cases, and for around half of them, we present in-body examples and user studies (Figure 5). From the sustainability perspective, all papers had a small statement declaring why they were better, ‘greener’ alternatives, four presented biodegradability studies [13, 14, 149] and two presented a schematic of their life cycle [82, 143].

Figure 6:

Figure 6: Materials used for different eTextile systems components.

In regard to the eTextile components, sensors are the most widely developed elements, and were mentioned in 14 sources (Figure 6). The papers specifically described sensors as suitable for measuring strain, moisture and pressure. The second most frequent purpose is to use biomaterials as conductive traces, either in yarn or textile format. However, signalling is only fabricated with carbon-based materials due to their electrical conductivity. Interestingly, power sources and supercapacitors (SC) were also demonstrated, contributing to more sustainable sources of energy generation and storage. Effectors such as light-transmitting materials and shape- and colour-changing textiles were also presented, together with RFID and antennae for wireless communication. Finally, substrates and cases were also shown as alternatives for commonly used plastics to support and store the previously mentioned components. Considering the perspective of eTextile systems, we found no examples for the use of biomaterial for eTextile microcontrollers.

4.2 Material Characteristics

The following section presents the found materials, providing information about their properties, formats, processes, and material combinations (Figure 7).

Figure 7:

Figure 7: Summary of biomaterials with their origin, main properties, formats, manipulation processes and combination with other materials.

4.2.1 Carbon-based materials.

Carbon-based materials are a diverse class of substances primarily composed of carbon atoms. They are known for their exceptional properties, including high strength, electrical conductivity, thermal stability, and biocompatibility. The versatility and adaptability of these materials have driven innovations across various fields, such as electronics, material science, and bio-medicine [12, 128]. Particularly related to the field of eTextiles, carbon-based materials have shown to be positively implemented in textiles in different ways [21, 22, 23, 24, 33, 88, 99, 146].

Within the array of these materials, graphene oxide (GO) has been used to create conductive silk fabric with fire retardant properties through dry-coating method, which consists of coating the fabric directly with a reduced GO hydrosol [68]. Spray-coating of GO onto facemasks has been used to create human respiration monitoring systems. The area coated with the conductive graphene works as a sensor, capturing the differences in resistance whenever the shape of the mask changes [17]. This type of strain sensor has also been developed by doing vacuum filtration of GO dispersion, where the cotton filter itself becomes the substrate of the sensor [118]. Modifying the surface of a sample can also be done by dip-coating a textile substrate into a tank containing the GO [136].

Another way to produce electrically active materials is carbonisation, a pyrolytic process that aims to convert a biomass product into a highly carbonaceous material. Liang et al. [2022] created supercapacitors by carbonising Bombyx mori silk. Molybdenum dioxide nanoparticles were used as a food additive for the silkworm. This resulted in modified silk yarns, achieving mass specific capacitance [89]. Choi et al. [2023] proposed another carbonisation product: a 3D fabric carbonised textile structure aimed to enhance the piezoelectric properties of a sensor. The 3D structure was fabricated by knitting the carbonised cotton into polyester, and then attaching it to a copper pressure sensor using silver paste [26].

In their versatility, carbon-based materials can also be used as dyes and inks. Liang et al. [2022] produced hydrophilic, biocompatible, and washable electronic textiles by simply dyeing the fabric in a sericin-graphene dispersion. This coating was then used as a strain sensor [89]. Akbari et al. [2016] fabricated textile-based antennae [1], by adding a graphene-based ink on the top of a cotton fabric substrate using the doctor blade technique [1, 4].

Carbon-nanotubes (CNT) can also be used in the creation of the fabric sensors [61]. The peculiarity of the process lies in the wrapping of CNT sheets around a yarn, either cotton or spandex, similarly to a spinning process. Then, to fabricate the biosensing electrodes, knitting, weaving and braiding were exploited [146].

Finally, carbon-based sweat wearables for energy harvesting have been developed using biofuel cells (BFV), and stored in an SC [92].

When working with carbon-based materials, however, it is important to keep in mind that one must work safely. The issue of their toxicity vs. biocompatibility is still under discussion and there are studies in the literature that lead to extremes of both possibilities [53]. The toxicity or biocompatibility of this class of materials seems to be closely related to their modification, size and concentration inside the human body upon inhaling or ingestion. That is why it is still an open question and fruit of various researches even nowadays [90, 97, 111, 113, 123].

Regarding end-of-life, it is known that the biodegradation of carbon-based materials is difficult, but there are studies that continue to refine the process using enzymes. Liu et al. [2022] reported how, thanks to the digestive action of microorganisms and enzymes in yellow mealworms´ guts, it is possible to biodegrade GO sheets [91]. In the case of Yang and Zhang [159], they focus their review on how the use of macrophages, white cells of the immune system specialised in the detection and elimination of harmful substances and bacteria, makes the biodegradation of different types of CNTs possible, pointing out that the time required for this process depends on different types of factors, such as their length and functional groups on their surface.

4.2.2 Cellulose.

Cellulose is a natural carbohydrate that serves as a primary structural component in the cell walls of plants and certain algae. It is one of the most abundant organic compounds on Earth and today plays a crucial role in industrial, technological and biomedical applications due to its mechanical properties, biocompatibility and easy modifiability. Cellulose, the source of which is plants, usually undergoes treatments to be extracted and made usable (i.e. nanofibrils and nanocrystals). These treatments, for instance the dissolution or mechanical separation of the plant fibres, lead to the presence of functional groups along the polymer chains.

Guridi et al. [2022] developed textile-based optical sensors using carboxymethylated cellulose (CMC). Two different types of CMC waveguides were achieved: fibres through wet spinning and planar by polymer casting. To create the optical sensors, both of the waveguides were weaved with either cotton or polyester yarns [50]. Guridi et al. [2023] also tested different weaving patterns and changed the ingredients of the waveguides, adding glycerine to enhance the flexibility of the material [51]. Wei et al. [2022] summarised the usage of cellulose as a ‘green triboelectric generator’. It can be implemented in various devices, such as pressure sensors, smart home systems, and medical and environmental monitoring systems [149].

Another studied source of nanocellulose comes not from plants, but from bacteria: bacterial cellulose. Commonly known as kombucha, it is a pure, highly crystalline form of cellulose, the repeating unit of which is the same as that found in plant cell walls. However, due to its purity and crystallinity bacterial cellulose is unique in its properties and production process [6]. Bell et al. [2023] have created an interactive breastplate using SCOBY (Symbiotic Culture of Bacteria and Yeast). SCOBY was grown in water, sugar and finally tea for the colour, which was added to the started culture. When it reached the adequate dimension, it was given proper shape and dried. Different electronic components can be embedded in it, such as control systems, LEDs and sensing parts [14]. Ng [2017] also worked with kombucha, similarly growing this bacterial nanocellulose, mixing it with dyes and embedding it with electronics [103]. Cellulose is also very well known to be biocompatible and biodegradable, like the above cited materials, but its degradation conditions and time depend on its different types of functionalisation (e.g. cellulose acetate, cellulose xanthate), creating different possibilities [35]. Hayakawa et al. [2014] have studied the effect of temperature on the rate of biodegradation, focusing on its burial in the forests of Japan, Indonesia and Thailand [58].

4.2.3 Polylactic acid (PLA).

Polylactic acid (PLA) is the biodegradable polymer obtained through the polycondensation of lactic acid, an organic acid that can be produced either via fermentation of, for example milk or whey, or chemical reactions. PLA is applied in a wide spectrum of fields, such as medical applications [122], agriculture [75], and automotive fields [105], and has recently attracted interest for the production of yarns for textiles [160]. PLA is now widely used in 3D printing and studied for its shape memory properties [157]. Leist et al. [2017] combined these two characteristics to create the 4D printing of PLA: a 3D model that can change shape when exposed to a heat source [86]. They first demonstrated how different shapes could be printed which, when compressed or stretched when hot, then cooled and brought back to 70 °C, permanently regain their original shape. This particular feature has been exploited to create a shape memory fabric in which nylon filaments are interwoven with 3D printed PLA.

Another example of the use of PLA for smart textiles is the work of Gong et al. [2022]. Their process design enabled them to obtain hollow PLA fibres that were filled with eutectic gallium-indium alloys, then weaved into a fabric to create reconfigurable green electronic textiles [45]. The main feature is that the two components, PLA and metal alloy, can be separated by a simple dissolution of the PLA fibres in dichloromethane, dissolving and recovering the gallium-indium alloy. Furthermore, Kalita et al. [2021] demonstrated the biodegradability of this material under different conditions and how it can be influenced by adding algae as additives [71].

4.2.4 Mycelium.

Mycelium is the vegetative, thread-like structure of fungi, consisting of a network of branching hyphae. It serves as the foundational part of the fungal organism and plays a crucial role in various ecological, industrial and technological contexts [5]. Like gelatine, mycelium, is often used as a stiffer scaffold to introduce or combine other functional materials or electric components. But, as a living system, the peculiarity of this natural material is that it can be grown directly in different desired shapes. Genç et al. [2022] have created different types of interactive samples. Crumbs of this fungus were placed inside polystyrene moulds and allowed to grow in desired shapes, and different components were added: LEDs for light-emitting devices, and heating plates for thermo-induced colour change or moving samples [41]. Similarly, Vasquez and Vega [2019] used mycelium to fabricate ‘Myco-accessories’. They also grew the material but used a lamination technique, sandwiching the electronic components between two thin layers of mycelium to create bracelets, crowns and other accessories. Mycelium is soaked into glycerine to confer more flexibility to the final piece [143]. Being a living system, it biodegrades once buried in soil, even when part of a composite. Gan et al. [2022] demonstrated this by burying their mycelium and bamboo samples in soil for two months, after which only the bamboo parts remained [40].

4.2.5 Gelatin.

Gelatin is a versatile protein derived from the collagen found in the connective tissues, bones and skins of animals, primarily those of cows and pigs. It has been used for centuries in various culinary, pharmaceutical and industrial applications due to its unique gelling, stabilising and binding properties. Moreover, unlike collagen, which is composed of three chains twisted together, gelatine is linear and easier to modify [2]. Gelatine, in combination with water and other components, can form strong, malleable or ductile foams, which are practical for extrusion or moulding to create desired shapes. Lazaro et al. [2022] created different types of interactive biofoam by exploring different fabrication techniques: moulding, layering, extrusion, and sewing. They produced colourful foams by adding dyes, conductive foam by adding steel fibres, and thermal and UV responsive accessories by adding photochromic or thermochromic pigments [82]. The biodegradation of gelatine, which can be derived from different animal sources, depends mainly on its molecular weight [121]. Martucci and Ruseckaite [2009] developed a three-layer based film with modified gelatine and compared its biodegradability with that of pure gelatine by burying both in soil: the latter showed a weight loss of 40% in one week [94].

4.2.6 Agar Agar.

Agar Agar, known also as Agar is a natural polysaccharide obtained from algae. Due to its special properties, it has been used for millennia in Asian cuisine and is widely known and used in many fields around the world. Agar agar is a versatile and valuable ingredient in the culinary and scientific industries because of its ability to form gels when combined with water [109].

Bell et al. [2022] explored the possibilities of their ‘Alganyl’, made simply by boiling agar together with glycerine and water; the mixture then casted and dried. Its re-cookability is an interesting aspect of this material, as it can be casted again for new purposes. As a film, Alganyl has been tested in different ways, such as weaving, knitting, pleating, and braiding. It can be mixed with dyes, and combined with other active materials. Conductive Alganyl is obtained by adding activated charcoal paste. Photo- and thermochromic features have also been achieved by mixing it with photoactive and thermoactive dyes. It can also be used to create 3D shapes, such as a purse [13].

Some gram-positive bacteria possess an enzyme called agarase, which hydrolyses agarose, one of the components of agar agar. In their work, Parashar and Kumar [2018], used some of these bacteria for agar degradation. An interesting result is that temperature influences the pH and action of agarase [110].

4.2.7 Chitosan.

Chitosan is a biodegradable and biocompatible polymer derived from chitin; a natural polymer found in the shells of crustaceans as well as in the cell walls of fungi. It is valued for its unique properties and has found a wide range of applications in various industries, such as agriculture, food, biomedics, and textiles [77]. Frequently, due to its good mechanical and antibacterial properties, chitosan is used as a biocompatible scaffold for making functional textiles. Tian et al. silver-coated polyamide with chitosan containing yarn through a spinning technique and weaved it into a cotton matrix. The resulting fabric as put in contact and separated with poly(dimethylsiloxane) (PDMS), which creates charges that are easily transportable by the silver. This results in a more eco-friendly tribogenerator [137]. The use and implementation of chitosan is a good biodegradable alternative to other plastics. Wronska et al. [2023] demonstrated this through the degradation of chitosan films in soil. Chitosan swells when exposed to moisture and thus allows microorganisms to penetrate it, aiding its biodegradation, which takes place within a few weeks [155].

4.2.8 Escherichia coli.

Escherichia coli (E. coli) is a species of Gram-negative bacteria present in both human and animal intestines. E. coli is a major topic of microbiological research and is used in efforts to improve public health because, while many strains are safe and helpful for digestion, others can infect people and lead to sickness [74]. Wang et al. [2017] focused on its application in humidity actuators by printing E. coli’s cell suspension onto cis-1,4-polyisoprene (latex) films. Moreover, through genetic engineering, they demonstrated the possibility of adding even new functionalities, in their case, fluorescence ability. Their work led to the creation of E. coli-based sweat-responsive wearables [148]. As a carbon-based living organism, E. coli eventually dies and biodegrades naturally. However, we found no papers on its actual biodegradation, probably because of a lack of interest in this field. Bacillus subtilis is a Gram-positive, rod-shaped bacterium that is widely recognised for its importance in various fields, including biology, biotechnology and agriculture. This bacterium has attracted attention for its unique characteristics and versatility, such as easy genetic manipulation and large-scale production [162]. Similarly to E. coli, it has been studied as an actuator because of its humidity-responsive properties. However, it has not been implemented in textile structures yet [161]. Also like E. coli, this living organism dies after some years and biodegrades naturally.

4.2.9 Quasi-solid ionic conductors.

Quasi-solid ionic conductors are a class of materials that exhibit both solid and liquid-like properties. These materials have attracted significant attention in various fields, especially as electrochemical devices and sensors, due to their ability to transport ions effectively while maintaining mechanical stability. Quasi-solid ionic conductors bridge the gap between traditional solid-state electrolytes and liquid electrolytes, offering unique advantages. The synthesis of quasi-solid ionic conductors can vary depending on the specific materials and applications involved. These materials often involve combining a polymeric or macromolecular matrix with a suitable ionic conductor, for example, electrolytes. Lei and Wu [2021] have described the potential of these materials. The dough created by mixing the polymer matrix with different salts is malleable and ductile. In their review and work they show that the addition of the salt improves the mechanical properties as well as the conductivity of these materials. It is even possible to create bio-based and edible dough, the application of which relies on the creation of smart skin, which acts as a mechanical sensor; this is possible by creating two layers of ionic conductive dough and separating them with an insulating material (e.g. insulating double-sided tape). The dough can even be 3D-printed or spun to create more unique structures [85].

Being made of gluten and starch, and being a protein and a polysaccharide, this dough can easily biodegrade. In their study, Domenek et al. [2004] tested the biodegradation of different gluten-based materials, showing that when buried, gluten takes around 50 days to completely degrade [29]. Furthermore, Jayasekara et al. [2003] demonstrated that in their starch-based film, all the starch degraded in 45 days, utilising an automated composting unit [67].

4.3 Interaction Possibilities

Figure 8:

Figure 8: Schematic representations of the materials, their formats and interaction goals. Orange colour represents the main biomaterials, and grey colour represents the supporting materials.

Figure 9:

Figure 9: Schematic representations of the materials, their formats and interaction goals. Orange colour represents the main biomaterials, and grey colour represents the supporting materials.

Figure 10:

Figure 10: Placing of material samples observed in the literature

Considering that eTextiles are meant for interaction, 20 publications clearly described how these human-biomaterial interactions could take place. The uses suggested included measuring external stimuli, physical activity, touch and pressure, or illuminating output elements (Figure 8 and 9). With one exception [86], the enabling eTextile components directly interact with the human body. Aligned with wearables, on-body placement is favoured, particularly on the neck, fingers, wrist, and forearm for motion-tracking prototypes (Figure 10). Positioned on the body, sample dimensions range from 2 to 20 cm [148], but only one paper describes an ambient display application [44].

To clarify their working principles, we present two categories:

(1)

Renewable and biodegradable biomaterials with similar properties to those found in commonly used eTextile materials (e.g. electrical conductivity), which enable interactions often encountered with smart textile interfaces.

(2)

Renewable and biodegradable biomaterials with new, different properties to those found in commonly used eTextile materials (e.g. hygromorphic), which enable new interactions with smart textile interfaces.

In the first case, carbon-based materials are predominantly employed to imbue textile substrates with electrical conductivity. The resulting fibres and textiles are then incorporated into conventional applications, such as bending sensors [26, 68, 89, 118, 137], strain sensors [118] or antennae [146]. Due to their renewable origin and capacity to biodegrade they offer a more sustainable option, as well as enabling already known applications and interactions.

In the latter case, biomaterials offer a more environmentally sustainable alternative, and also enhance potential interactions within eTextile systems through their intrinsic properties. Active materials such as cellulose, PLA, bacteria, and agar agar dynamically respond to environmental changes, facilitating shape deformations in response to humidity [86, 148] or variations in optical qualities [50]. These transformative attributes distinguish them from conventional eTextile materials and enable novel interactions such as the ability to perceive haptic changes through bacteria moving wearables [148]. In addition, mycelium, agar agar and bacterial cellulose are capable of ‘growing’ and encapsulating diverse functional components, such as thermochromic inks or microcontrollers, creating new mediating interfaces between functional materials and the human body.

Notably, these bio-based interfaces introduce distinctive characteristics in comparison to the traditional use of plastics or metals in eTextiles. The tangible qualities such as agar agar´s transparency [13], the natural feeling of mycelium [41], and the soft and flexible texture of gelatine [82] introduce interesting tactile interaction possibilities, such as the ability to press or squeeze or establish an intimate connection between a ‘lifelike’ bacterial cellulose and the human body [14].

Conclusively, their versatile manipulation capabilities, including coating, foaming, moulding, cutting, weaving, or knitting (Table 7), underscore the potential of these materials to craft a diverse array of eTextile systems — ranging from necklaces and headbands to breastplates. Again, this amplifies the potential for curated interactions among functional materials, the environment, and the human body.

In order to better understand the potential of the new interactions enabled by these biomaterials, authors [13, 14, 41] delved into the experiential qualities of biomaterial-based eTextiles. Their main insights highlight that in comparison to traditional eTextile materials, biomaterials can be perceive as fragile, provoke a feeling of ‘uncanniness’ or ‘creepiness’ [41]. They can also provide sensory-rich experiences [82], foster a sense of ‘shared livingness’ between the human wearer and the non-human organism [14], and present positively charged characteristics such as biodegradability, flexibility, and ease of fabrication [13].

Skip 5INSIGHTS FROM THE LITERATURE Section

5 INSIGHTS FROM THE LITERATURE

The preceding results section highlighted 9 biomaterials identified in contemporary research for constructing components within eTextile systems. It presented their key attributes, including biodegradability and biocompatibility, along with their diverse array of applications and formats (Figures 7, 8, 9). A closer examination of the literature unveiled a promising avenue for developing sustainable eTextiles, in which the inherent qualities of bio-based materials enable novel interactions. These materials can actively respond to the environment, integrate functional components, undergo biodegradation, and be manipulated through various techniques such as weaving, knitting, printing, moulding, growing, pleating, cutting, or sewing. The next section presents an in-depth analysis of the results, and offers seven primary insights:

(1)

Biomaterials for eTextiles are an under-explored field

(2)

Fully bio-based eTextile systems can be constructed

(3)

Biodegradability is a design variable

(4)

Even though promising, the environmental impact of biomaterials for eTextiles remains unclear

(5)

DIY approaches are contributing to material innovation in the interdisciplinary field of eTextiles

(6)

Biomaterials provide an opportunity to amplify the interactive possibilities of eTextiles

(7)

The aesthetics and interpretative understandings of biomaterials for eTextiles should be deepened

5.1 Biomaterials for eTextiles Are an Under-explored Field

The first notable issue emerging from the literature is a clear research gap in the use of biomaterials for eTextiles. We found only 33 publications in 19 different outlets, including various journals and conferences.

Half of these sources revolve around carbon-based materials, a well-established domain within novel conductive materials research. Their lightweight, robust, and conductive properties render them ideal for crafting various components within eTextile systems, including sensors and antennae. Nevertheless, the number of publications explicitly addressing the reasons and methodologies for utilising these materials as more sustainable alternatives is notably insufficient in comparison to the overall findings. The other half of the results comprise height distinct material categories: mycelium, cellulose, gelatine, salts, agar agar, chitosan, PLA, and quasi-solid ionic conductors. Each category is underrepresented, with only one to three papers discovered for each one, all published between 2017 and 2023. This scarcity highlights that this research area is still in its initial stages.

Regarding knowledge clusters, material science is predominant, followed by HCI and electrical engineering. In contrast, textile design has minimal representation, and we found only two papers on it. This supposes a relevant research gap to be addressed by future research. Knowledge steaming from textile research and practice is fundamental to enrich the field of sustainable eTextiles. Integrating deeper understandings on textile thinking theory [62], fabrication methods that could enhance functionalities [9, 115], or circularity for fashion and textiles [104, 141], among others, will contribute to more holistic and purposeful developments of biobased eTextiles. Additionally, insights from the fields of green electronics [132] and industrial design [70], should be integrated through dedicated literature reviews of biomaterials’ applications within these domains. It is relevant that these contributions stem from interdisciplinary teams, as their diverse methodologies, resources and epistemologies could enrich the findings. Collaboration should be actively encouraged through initiatives such as the Bioinnovation Center [140], Software [126], STARTS Programme [117], or Bio-Inspired Textiles [63].

5.2 Fully Bio-based eTextiles Systems Can Be Constructed

Understanding eTextiles as systems highlighted the versatility of biomaterials for creating different components, taking into account not only conductive elements but also those that can contain, bind, protect, react, or store. The catalogue of biomaterials we found presents an array of different formats and processes, each offering multiple possibilities, as shown in Figure 7. Carbon-based materials such as inks, filaments, pastes and liquid solutions showed potential to create various elements: power sources, conductive traces, sensors, effectors, passive elements, antennae, and RFID tags. In addition, the use of conductive salts in dough shapes enable the creation of flexible and malleable sensors. Mycelium and gelatine exhibit great potential for encapsulating other materials, thereby enhancing their capabilities as sensors and effectors. Simultaneously, they take on the form of malleable 3D shapes, serving as both substrates and protective casings. Agar agar, presented in its film configuration, can undergo diverse transformations through techniques such as heat sealing, weaving, knitting, pleating, braiding, and laser engraving, leading to the creation of colourful, flexible substrates. In the case of Escherichia coli, living bacteria enabled textile movement, avoiding the need for metallic or plastic fibres. Lastly, cellulose, whether derived from plants or bacteria, proved to be useful for power sources, sensors, effectors, and substrates. This is particularly interesting due to the growing interest in green electronics and the use of cellulose for a wide range of components such as substrates for flexible electronics [66] or strain sensors [11].

The integration of these biomaterials offers boundless opportunities to replace nearly all components in an eTextile system. Components made from different biomaterials could be combined in the same systems, or go even further and propose mono-material systems (e.g. with cellulose). No examples of bio-based microcontrollers were found in the literature, but inspiration could be taken from non-traditional computer constructions such as the 8-bit embroidery computer made by Posh and Kurbak [2018] [64]. This potential could be further magnified by embracing modularity [55] and disassembly [39, 156] as integral design principles, facilitating material separation and component repair at the end-of-life stages. Consequently, we envision biomaterials as central agents in the realisation of fully bio-based eTextile systems, moving toward sustainable and eco-friendly textile technology.

5.3 Biodegradability Is a Design Variable

From the perspective of design, we considered biomaterials as coming from renewable biomass origins and/or as having the capacity to biodegrade. As mentioned in the definitions section, biodegradability means that the materials can be degraded through microorganisms available in the environment, converting them into natural substances such as water, carbon dioxide and composts. This characteristic can greatly contribute to the end-of-life of eTextile products as part of circular processes in which materials can return to the soil without throwbacks. To determine the actual impact of the biodegradability and composting of the materials, collection and disposal systems need to be taken into consideration. In some cases, the composting of biomaterials can bring more challenges, due to the lack of special facilities for textile-related biomaterials, manipulation processes and the release of toxic gases [119]. In the reviewed literature, biodegradation appears to be one of the main material beneficial characteristics. The agar agar user’s study revealed that biodegradability was the most interesting aspect for the participants, inspiring them to imagine future uses of Alganyl [13]. Three publications described biodegradability studies. One prototype of the cellulose-based energy storage element [127] was dug up and only 30% of Vim’s mass remained after 60 days. Bell et al. [2023] buried a 5 cm by 5 cm SCOBY sample in soil containing living microbes, which degraded by 96% in 30 days [14]. In the case of agar agar [13], a paper claimed that Alganyl degraded by 97% in 60 days in a controlled environment kept at 40° C.

Biodegradability is thus becoming an important variable to consider when designing eTextile systems, as it can contribute a closer loop, either by returning to the soil, enabling disassembly or reutilisation. In addition, we see how degradability can also contribute to the material aesthetics and affordances of interactive systems. The time transformation of the material, due to interaction with people and the environment, could produce shape- or colour-shifting materials, opening an exploration path for novel interactions, as for example, in the artistic work of ‘Cambio de Piel’ [48], in which a text is revealed on a fabric layer as the top biomaterial layer dissolves in water. In the case of wearables design, these properties can also contribute to customisation and the exploration of ephemeral fashion [81]. This changeability approach has been proposed by Talman [2019], and offers new ways of exploring transformations in colour, texture and structure in an aim to understand how we can encourage acceptance of changes in textiles and re-establish textile–people connections [135].

5.4 Environmental Impact of Biomaterials for eTextiles Remains Unclear

The results presented offer an inspiring justification of the exploration of more sustainable options in eTextile creation, driven by the advantages of biomaterials – their renewability, biodegradability, and, in most cases, biocompatibility. However, this is still a young research area, with all the material outcomes falling under Technology Readiness Level 1 [101]. At this stage, the research primarily revolves around establishing fundamental principles and formulating initial technology concepts and applications. Out of the identified publications, only five presented a biodegradability study [13, 14, 99], and two presented a life cycle schematic [82, 143]. None of them presented a comprehensive life cycle assessment. This indicates the lack of clarity regarding the environmental impact of these materials. Ideally, a circular perspective should be ingrained from TRL1 onwards, as advocated by Schinchke et al. [2020]. It is crucial to continually reflect on the possibilities and challenges associated with scaling up the production of these materials, including their influence and interconnections with various factors of the life cycle [142], such as transportation, design, usage, and end-of-life considerations. Interdisciplinary collaboration is crucial in this objective, as it facilitates an inclusive analysis that incorporates laboratory-based investigations alongside social and economic factors from a systemic perspective [31]. Critical perspectives are vital for examining production levels, material origins, treatment processes, and the involvement of communities within their ecosystems [150]. For insights into addressing production challenges and understanding their implications on a larger scale, business experiences from eTextile [125] and biomaterials such as MycoWorks [100] can be instrumental. Lastly, staying attuned to the evolving legislation concerning Bioplastics and Biomaterials [36] will also provide guidance for future directions and unveil potential investment opportunities in this field.

5.5 DIY Approaches Contributing to Material Innovation

We identified a total of 30 distinct manufacturing processes, with the possibility of applying multiple methods to each material, as detailed in Table 7. Carbon-based materials often necessitate specialised laboratory facilities and equipment, such as vacuum filtration [118] and carbonisation [26]. Scientific research frequently uses nanoparticles that macroscopically result in a powder, requiring the usage of proper safety gear. More user-friendly suspensions are available for commercial applications, lessening the risk of safety hazards. Despite their format, some of these materials are often chemically modified to improve their chemical or physical properties, depending on the application; one common example is the chemical reduction of GO to enhance its electrical properties [65].

For E. coli [148], manipulation took place in special biolabs under a controlled environment that guaranteed biosafety, and had to obtain security approval. Even if naturally present in our intestines, some strains of E. coli can cause severe infections. In contrast, materials such as mycelium, gelatine, cellulose, chitosan, agar agar, and various salts are more accessible for purchase and easier to manipulate. The associated processes are closely related to cooking and textile techniques, eliminating the need for specialised laboratory facilities. They can be executed in a do-it-yourself (DIY) manner utilising open-source resources from various repositories and educational programmes such as Materiom [95], Fabricademy [37], Healthy Materials Lab [79], Chemarts [70], and Biodesign Challenge [25]. This combination of DIY techniques and resource networks enables non-scientists to more freely prototype and try different fabrication such as mould casting [13], weaving [51] or laser cutting [143]. Modifying moulds, combining materials, and altering scale, colours, or textures further enables the discovery of new material behaviours.

The literature review also highlights the potential for these materials to serve as playgrounds for affordance studies and collaborative workshops, inviting non-experts to engage in material manipulation and ideation processes [13]. Therefore, we see how these material ‘playgrounds’ [16] can open the path for innovative processes and material combinations, which can be guided through design methods such as fast prototyping and material-driven design frameworks [73]. This can be particularly valuable in the case of interdisciplinary collaboration, allowing designers to lower the entry barriers to understanding new materials through hands-on work, and for scientists to approach the material in different settings, new scales and from a more open-ended ideation process. Designers benefit by gaining a practical understanding of new materials through hands-on work, and scientists get to explore materials in diverse settings, at different scales, and through more open-ended ideation processes [13, 116]. In conclusion, these DIY approaches offer the potential to discover innovative processes and material combinations, simplifying the integration of new materials into design practices and encouraging more versatile scientific exploration.

5.6 Amplifying Interaction Posibilities

The analysis of interactivity (Section 4.3) revealed that biomaterials can significantly contribute to advancing current developments in soft interactive interfaces within HCI. They can, for example, provide carbon-based conductive yarns for smart sanitary napkins [98] in order to contribute to their disposal and biodegradability challenges, or present new bio-based tactile qualities to keep exploring affordances of surface gestures on textile user interfaces [96].

These biomaterials not only serve as substitutes for traditional materials but also enhance the interactive capabilities of eTextile systems in HCI. Beyond offering a more sustainable alternative to conventional plastics and metals, these materials introduce novel properties that foster the exploration of new relationships among humans, the environment and interfaces. The diverse textures, such as rugged mycelium or adhesive biofoam, alongside various formats such as cellulose films or PLA volumes provide researchers in the HCI field with myriad options for experimenting with sustainable materials, thereby supporting the creation of multisensorial interactive experiences. A distinctive feature of biomaterials lies in their being both biodegradable and active materials, making them unique in their transformative capacity. These materials can deform, distort, shrink, swell, or degrade over time and in response to different environmental stimuli, introducing temporality as a new design variable [139]. This characteristic brings forth the concepts of ‘growth’ and ‘liveness,’ contributing to ongoing discussions on more-than-human entities and their agency in design processes and interactions.

An additional insight obtained from the literature pertains to the scale of eTextile examples predominantly situated on the human body in direct contact with the skin. This inclination can be attributed to the intrinsic relevance of eTextiles within the wearable technology domain, to which the majority of applications are targeted. To maintain this on-body focus, the scale of material samples and prototypes spans a spectrum ranging from a mere 1 cm in the case of conductive yarns and prints [30] to approximately 30 cm for the humidity-reactive shirt [148]. As an exception, the mycelium study [41] presents a unique potential for application in interior design elements and ambient displays. We acknowledge the relevance and importance of small-scale, but we also see an unexplored field of larger bio-based interactive interfaces.

Given the versatility of biomaterial production techniques and their accessibility, we advocate for the expansion of sample dimensions. As biomaterial-based eTextiles do not use scarce resources, heavy metals or toxic chemicals, which urge minimising electrically functional material consumption, this opens up the exciting prospect of creating larger interactive interfaces that may not necessarily adhere to the body but can facilitate novel HCIs through these materials. Examples of larger-scale interaction can be found in interaction design, such as electrostatic interior textiles [47], which involve room-size textile installations designed for triboelectric energy generation capable of shaping our everyday behaviours. As the electrostatic textiles did not specifically focus on biomaterials, research on piezoelectric biomaterials [147] could provide directions for further research. Furthermore, for more extensive biomaterial surfaces, 1.5-metre tapestry pieces of hand-woven cellulose have been crafted for an interactive installation called ‘Borrowed Matter’ [49]. The potential for upscaling material interfaces broadens the horizon for novel applications, exposing them to wider audiences and fostering collaboration with artists, architects, stage designers, performers, and construction engineers. These explorations could expand the creative landscape of the field of human-computer interaction.

In summary, biomaterials emerged as catalysts for expanding possibilities within HCI, not only through their environmental sustainability but also through their dynamic and transformative nature, fostering a realm of novel interactive experiences.

5.7 Deepening Aesthetics and Interpretative Understandings

Understanding how users attribute value to specific eTextile products holds the potential to foster sustainable consumption habits and encourage long-lasting usage [108]. In this context, the aesthetics of materials assume an important role in shaping users’ perceptions and experiences of interactive systems. In this regard, five publications [13, 14, 41, 51, 82] explicitly manifested a keen interest in the aesthetics of their prototypes. Fiona Bell et al. [2023] shared their personal reflections regarding their immersive experience with the SCOBY breastplate, simultaneously acknowledging and embracing the material’s ‘imperfections’ [14]. The natural colours and organic textures were regarded as personal and connecting elements that forged a bond between the practitioners and the material. Meanwhile, Guridi et al. [2023], in their creation of cellulose-based optical sensors, detailed their decision-making process concerning complementary materials and colour palettes [51]. They accompanied this with comprehensive visual documentation aimed at enriching the evocative visual narratives surrounding the material. Extending this comprehension, user studies were presented in the cases of mycelium [41] and agar agar [13]. In the first case, the quest for experiential qualities led researchers to ask for participant feedback on whether the dynamic changes in mycelium samples were ‘aesthetically pleasing’ and to explore the meanings and emotions elicited by the material. These findings revealed ideas on future applications and design directions for interactive mycelium products. In the case of agar agar, a user study served as a lens through which to analyse the material’s desirable attributes and drawbacks, thereby identifying prospective applications [13]. The authors also extensively explored colours, shapes and textures.

We observe that integrating these qualitative studies into material development could inspire practitioners and researchers to explore new possibilities, harnessing user experiences and interpretations to craft more meaningful and purpose-driven eTextile products [152]. Material-driven design methodologies [73] offer a framework for conducting these studies and facilitating the effective communication of results within interdisciplinary teams.

Skip 6CONCLUSION Section

6 CONCLUSION

Our comprehensive literature review on Biomaterials for eTextiles highlighted sustainable options for developing interactive textile interfaces. Despite the limited existing research, this effort synthesised insights from different fields such as HCI, Material Science, Electrical Engineering, and Textiles. To further advance this research, we suggest incorporating more knowledge steaming from Textile research, Green Electronics, and Design, even if not directly eTextile-related.

The findings unveil promising directions, including the potential creation of fully bio-based eTextile systems. In this context, biomaterials emerge as increasingly sustainable alternatives, simultaneously enhancing interaction possibilities due to their distinctive characteristics. To comprehensively grasp their potential, integrating thorough life cycle assessments is imperative for evaluating their authentic environmental footprint.

Finally, we claim that these advancements necessitate interdisciplinary collaboration, integrating scientific methodologies, user studies, eco-design guidelines, and aesthetic considerations. Such a holistic approach is key to unveiling the full potential of biomaterials in crafting sustainable and innovative eTextile solutions.

Skip Supplemental Material Section

Supplemental Material

Video Preview

Video Preview

mp4

44.8 MB

Video Presentation

Video Presentation

mp4

22.2 MB

References

  1. M Akbari, J Virkki, L Sydänheimo, and L Ukkonen. 2016. Toward graphene-based passive UHF RFID textile tags: A reliability study. IEEE Transactions on Device and Materials Reliability 16, 3 (2016), 429–431.Google ScholarGoogle ScholarCross RefCross Ref
  2. J Alipal, NAS Mohd Pu’Ad, TC Lee, NHM Nayan, N Sahari, H Basri, MI Idris, and HZ Abdullah. 2021. A review of gelatin: Properties, sources, process, applications, and commercialisation. Materials Today: Proceedings 42 (2021), 240–250.Google ScholarGoogle ScholarCross RefCross Ref
  3. Ananas Anam. 2023. Piñatex. Retrieved September, 2023 from https://www.ananas-anam.com/sales-sampling/Google ScholarGoogle Scholar
  4. Grigoria Athanasaki, Arunkumar Jayakumar, and AM Kannan. 2023. Gas diffusion layers for PEM fuel cells: Materials, properties and manufacturing–A review. International Journal of Hydrogen Energy 48, 6 (2023), 2294–2313.Google ScholarGoogle ScholarCross RefCross Ref
  5. Noam Attias, Ofer Danai, Tiffany Abitbol, Ezri Tarazi, Nirit Ezov, Idan Pereman, and Yasha J Grobman. 2020. Mycelium bio-composites in industrial design and architecture: Comparative review and experimental analysis. Journal of Cleaner Production 246 (2020), 119037.Google ScholarGoogle ScholarCross RefCross Ref
  6. Tariq Aziz, Arshad Farid, Fazal Haq, Mehwish Kiran, Asmat Ullah, Kechun Zhang, Cheng Li, Shakira Ghazanfar, Hongyue Sun, Roh Ullah, 2022. A review on the modification of cellulose and its applications. Polymers 14, 15 (2022), 3206.Google ScholarGoogle ScholarCross RefCross Ref
  7. Kees Baldé, E D’Angelo, V Luda, O Deubzer, and Rüdiger Kühr. 2022. Global transboundary e-waste flows monitor 2022. United Nations Institute for Training and Research (UNITAR).Google ScholarGoogle Scholar
  8. Hossein Baniasadi, Zahra Madani, Mithila Mohan, Maija Vaara, Sami Lipponen, Jaana Vapaavuori, and Jukka V Seppala. 2023. Heat-Induced Actuator Fibers: Starch-Containing Biopolyamide Composites for Functional Textiles. ACS Applied Materials & Interfaces 15, 41 (2023), 48584–48600.Google ScholarGoogle ScholarCross RefCross Ref
  9. HuiYing Bao, Yan Hong, Tao Yan, Xiufen Xie, and Xianyi Zeng. 2023. A systematic review of biodegradable materials in the textile and apparel industry. The Journal of The Textile Institute (2023), 1–20.Google ScholarGoogle ScholarCross RefCross Ref
  10. Bahareh Barati, Elvin Karana, Sylvia Pont, and Tim Van Dortmont. 2021. LIVING LIGHT INTERFACES—AN EXPLORATION OF BIOLUMINESCENCE AESTHETICS. In Designing Interactive Systems Conference 2021. 1215–1229.Google ScholarGoogle ScholarDigital LibraryDigital Library
  11. Fevzihan Basarir, Joice Jaqueline Kaschuk, and Jaana Vapaavuori. 2022. Perspective about cellulose-based pressure and strain sensors for human motion detection. Biosensors 12, 4 (2022), 187.Google ScholarGoogle ScholarCross RefCross Ref
  12. Aneeqa Bashir, Azka Mehvish, and Maria Khalil. 2021. Advanced carbon materials for sustainable and emerging applications. 21st century advanced carbon materials for engineering applications-A comprehensive handbook (2021).Google ScholarGoogle Scholar
  13. Fiona Bell, Latifa Al Naimi, Ella McQuaid, and Mirela Alistar. 2022. Designing with Alganyl. In Sixteenth International Conference on Tangible, Embedded, and Embodied Interaction. 1–14.Google ScholarGoogle Scholar
  14. Fiona Bell, Derrek Chow, Hyelin Choi, and Mirela Alistar. 2023. SCOBY Breastplate: Slowly Growing a Microbial Interface. In Proceedings of the Seventeenth International Conference on Tangible, Embedded, and Embodied Interaction. 1–15.Google ScholarGoogle ScholarDigital LibraryDigital Library
  15. Fiona Bell, Netta Ofer, and Mirela Alistar. 2022. ReClaym our Compost: Biodegradable Clay for Intimate Making. In Proceedings of the 2022 CHI Conference on Human Factors in Computing Systems. 1–15.Google ScholarGoogle ScholarDigital LibraryDigital Library
  16. Fiona Bell, Netta Ofer, Ethan Frier, Ella McQuaid, Hyelin Choi, and Mirela Alistar. 2022. Biomaterial Playground: Engaging with Bio-based Materiality. In CHI Conference on Human Factors in Computing Systems Extended Abstracts. 1–5.Google ScholarGoogle Scholar
  17. Hossein Cheraghi Bidsorkhi, Negin Faramarzi, Babar Ali, Lavanya Rani Ballam, Alessandro Giuseppe D’Aloia, Alessio Tamburrano, and Maria Sabrina Sarto. 2023. Wearable Graphene-based smart face mask for Real-Time human respiration monitoring. Materials & Design 230 (2023), 111970.Google ScholarGoogle ScholarCross RefCross Ref
  18. European Bioplastics. 2023. European Bioplastics. Retrieved September, 2023 from https://www.european-bioplastics.orgGoogle ScholarGoogle Scholar
  19. Leah Buechley and Ruby Ta. 2023. 3D Printable Play-Dough: New Biodegradable Materials and Creative Possibilities for Digital Fabrication. In Proceedings of the 2023 CHI Conference on Human Factors in Computing Systems. 1–15.Google ScholarGoogle ScholarDigital LibraryDigital Library
  20. Inés Macarena Burdiles Araneda, Xavier Dominguez, Marion Real, Santiago Fuentemilla, and Anastasia Pistofidou. 2022. Remix The School Project: Socio-Emotional Learning through Biomaterial making: A Methodology for Self-Awareness through Bio-material fabrication for teachers and children. In 6th FabLearn Europe/MakeEd Conference 2022. 1–5.Google ScholarGoogle ScholarDigital LibraryDigital Library
  21. Haihua Cai, Yujia Wang, Mengting Xu, Lan Cheng, Zulan Liu, Zhi Li, and Fangyin Dai. 2022. Low cost, green and effective preparation of multifunctional flexible silk fabric electrode with ultra-high capacitance retention. Carbon 188 (2022), 197–208.Google ScholarGoogle ScholarCross RefCross Ref
  22. José Tiago Carvalho, Inês Cunha, João Coelho, Elvira Fortunato, Rodrigo Martins, and Luís Pereira. 2022. Carbon-yarn-based supercapacitors with in situ regenerated cellulose hydrogel for sustainable wearable electronics. ACS Applied Energy Materials 5, 10 (2022), 11987–11996.Google ScholarGoogle ScholarCross RefCross Ref
  23. Pietro Cataldi, Marco Cassinelli, José A Heredia-Guerrero, Susana Guzman-Puyol, Sara Naderizadeh, Athanassia Athanassiou, and Mario Caironi. 2020. Green biocomposites for thermoelectric wearable applications. Advanced Functional Materials 30, 3 (2020), 1907301.Google ScholarGoogle ScholarCross RefCross Ref
  24. Pietro Cataldi, Pietro Steiner, Mufeng Liu, Gergo Pinter, Athanassia Athanassiou, Coskun Kocabas, Ian A Kinloch, and Mark A Bissett. 2023. A Green Electrically Conductive Textile with Tunable Piezoresistivity and Transiency. Advanced Functional Materials (2023), 2301542.Google ScholarGoogle Scholar
  25. Biodesign Challenge. 2024.. Retrieved January, 2024 from https://www.biodesignchallenge.org/Google ScholarGoogle Scholar
  26. Hyeongsub Choi, Jingzhe Sun, Bingqi Ren, Seokjun Cha, Jiwoo Lee, Byoung-Min Lee, Jin-Ju Park, Jae-Hak Choi, and Jong-Jin Park. 2022. 3D textile structure-induced local strain for a highly amplified piezoresistive performance of carbonized cellulose fabric based pressure sensor for human healthcare monitoring. Chemical Engineering Journal 450 (2022), 138193.Google ScholarGoogle ScholarCross RefCross Ref
  27. Daniel Corzo, Guillermo Tostado-Blázquez, and Derya Baran. 2020. Flexible electronics: status, challenges and opportunities. Frontiers in Electronics 1 (2020), 594003.Google ScholarGoogle ScholarCross RefCross Ref
  28. EC Directive. 2012. Directive 2012/19/EU of the European Parliament and of the Council of 4 July 2012 on waste electrical and electronic equipment, WEEE. Official Journal of the European Union L 197 (2012), 38–71.Google ScholarGoogle Scholar
  29. Sandra Domenek, Pierre Feuilloley, Jean Gratraud, Marie-Hélène Morel, and Stéphane Guilbert. 2004. Biodegradability of wheat gluten based bioplastics. Chemosphere 54, 4 (2004), 551–559.Google ScholarGoogle ScholarCross RefCross Ref
  30. Yongchun Dong, Xuan Sun, and Jiayu Gu. 2023. Surface carboxylation of hydrophobic synthetic fibers for enhancing deposition of reduced graphene oxide to create highly conductive and bactericidal textiles. The Journal of The Textile Institute 114, 8 (2023), 1135–1145.Google ScholarGoogle ScholarCross RefCross Ref
  31. S Duarte Poblete, Laura Anselmi, Valentina Rognoli, 2023. Emerging materials fostering interdisciplinary collaboration in Materials Design. In Interdisciplinary Practice in Industrial Design. AHFE Open Access, 119–129.Google ScholarGoogle Scholar
  32. Marzia Dulal, Shaila Afroj, Jaewan Ahn, Yujang Cho, Chris Carr, Il-Doo Kim, and Nazmul Karim. 2022. Toward sustainable wearable electronic textiles. ACS nano 16, 12 (2022), 19755–19788.Google ScholarGoogle Scholar
  33. Bel Hadj Jrad Elyes, Chebil Achref, and Dridi Chérif. 2022. Biomass derived quasi solid state supercapacitors for smart textile integration. In 2022 IEEE International Conference on Design & Test of Integrated Micro & Nano-Systems (DTS). IEEE, 1–6.Google ScholarGoogle Scholar
  34. Elisabeth Eppinger, Alina Slomkowski, Tanita Behrendt, Sigrid Rotzler, and Max Marwede. 2022. Design for Recycling of E-Textiles: Current Issues of Recycling of Products Combining Electronics and Textiles and Implications for a Circular Design Approach. In Recycling-Recent Advances. IntechOpen.Google ScholarGoogle Scholar
  35. Nejla B Erdal and Minna Hakkarainen. 2022. Degradation of cellulose derivatives in laboratory, man-made, and natural environments. Biomacromolecules 23, 7 (2022), 2713–2729.Google ScholarGoogle ScholarCross RefCross Ref
  36. EU. 2023. BioplasticsEurope. Retrieved September, 2023 from https://bioplasticseurope.eu/projectGoogle ScholarGoogle Scholar
  37. Fabricademy. 2024. Fabricademy Tutorials. Retrieved January, 2024 from https://class.textile-academy.org/tutorials/Google ScholarGoogle Scholar
  38. Ylva Fernaeus and Petra Sundström. 2012. The material move how materials matter in interaction design research. In proceedings of the designing interactive systems conference. 486–495.Google ScholarGoogle ScholarDigital LibraryDigital Library
  39. Laetitia Forst. 2020. Disassembly discussed: creative textile sampling as a driver for innovation in the circular economy. Journal of Textile Design Research and Practice 8, 2 (2020), 172–192.Google ScholarGoogle ScholarCross RefCross Ref
  40. Jun Ken Gan, Eugene Soh, Nazanin Saeidi, Alireza Javadian, Dirk E Hebel, and Hortense Le Ferrand. 2022. Temporal characterization of biocycles of mycelium-bound composites made from bamboo and Pleurotus ostreatus for indoor usage. Scientific Reports 12, 1 (2022), 19362.Google ScholarGoogle ScholarCross RefCross Ref
  41. Çağlar Genç, Emilia Launne, and Jonna Häkkilä. 2022. Interactive Mycelium Composites: Material Exploration on Combining Mushroom with Off-the-shelf Electronic Components. In Nordic Human-Computer Interaction Conference. 1–12.Google ScholarGoogle Scholar
  42. Roland Geyer, Jenna R Jambeck, and Kara Lavender Law. 2017. Production, use, and fate of all plastics ever made. Science advances 3, 7 (2017), e1700782.Google ScholarGoogle Scholar
  43. Claudio Gioia, Greta Giacobazzi, Micaela Vannini, Grazia Totaro, Laura Sisti, Martino Colonna, Paola Marchese, and Annamaria Celli. 2021. End of life of biodegradable plastics: composting versus Re/upcycling. ChemSusChem 14, 19 (2021), 4167–4175.Google ScholarGoogle ScholarCross RefCross Ref
  44. Gozde Goncu-Berk. 2019. Smart textiles and clothing: An opportunity or a threat for sustainability. Proceedings of the Textile Intersections (2019).Google ScholarGoogle Scholar
  45. Wei Gong, Yang Guo, Weifeng Yang, Zhihua Wu, Ruizhe Xing, Jin Liu, Wei Wei, Jie Zhou, Yinben Guo, Kerui Li, 2022. Scalable and reconfigurable green electronic textiles with personalized comfort management. ACS nano 16, 8 (2022), 12635–12644.Google ScholarGoogle Scholar
  46. Phillip Gough, Praneeth Bimsara Perera, Michael A Kertesz, and Anusha Withana. 2023. Design, Mould, Grow!: A Fabrication Pipeline for Growing 3D Designs Using Myco-Materials. In Proceedings of the 2023 CHI Conference on Human Factors in Computing Systems. 1–15.Google ScholarGoogle ScholarDigital LibraryDigital Library
  47. Ramyah Gowrishankar. 2020. Engaging with e-static textiles: An investigation into textile interaction design for shaping body-driven energy harvesting in the interior. Ph. D. Dissertation. Universität der Künste Berlin.Google ScholarGoogle Scholar
  48. Sofia Guridi. 2022. Change of skin. Retrieved September, 2023 from https://sofiaguridi.xyz/change-of-skinGoogle ScholarGoogle Scholar
  49. Sofia Guridi. 2023. Borrowed Matter. Retrieved September, 2023 from https://www.revistamateria.com/borrowedmatter/Google ScholarGoogle Scholar
  50. Sofía Guridi, Ari Hokkanen, Aayush Kumar Jaiswal, Nonappa Nonappa, and Pirjo Kääriäinen. 2022. Integration of carboxymethyl cellulose waveguides for smart textile optical sensors. In 2022 IEEE Sensors. IEEE, 1–4.Google ScholarGoogle Scholar
  51. Sofía Guridi, Emmi Pouta, Ari Hokkanen, and Aayush Jaiswal. 2023. LIGHT TISSUE: Development of cellulose-based optical textile sensors. In Proceedings of the Seventeenth International Conference on Tangible, Embedded, and Embodied Interaction. 1–14.Google ScholarGoogle ScholarDigital LibraryDigital Library
  52. Olga Gurova, Timothy Robert Merritt, Eleftherios Papachristos, and Jenna Vaajakari. 2020. Sustainable solutions for wearable technologies: Mapping the product development life cycle. Sustainability 12, 20 (2020), 8444.Google ScholarGoogle ScholarCross RefCross Ref
  53. Sangiliyandi Gurunathan and Jin-Hoi Kim. 2016. Synthesis, toxicity, biocompatibility, and biomedical applications of graphene and graphene-related materials. International journal of nanomedicine (2016), 1927–1945.Google ScholarGoogle Scholar
  54. Lon Åke Erni Johannes Hansson, Teresa Cerratto Pargman, and Daniel Sapiens Pargman. 2021. A decade of sustainable HCI: connecting SHCI to the sustainable development goals. In Proceedings of the 2021 CHI Conference on Human Factors in Computing Systems. 1–19.Google ScholarGoogle ScholarDigital LibraryDigital Library
  55. Dorothy Hardy, Rachael Wickenden, and Angharad McLaren. 2020. Electronic textile reparability. Journal of Cleaner Production 276 (2020), 124328.Google ScholarGoogle ScholarCross RefCross Ref
  56. Dorothy A Hardy, Andrea Moneta, Viktorija Sakalyte, Lauren Connolly, Arash Shahidi, and Theodore Hughes-Riley. 2018. Engineering a costume for performance using illuminated LED-yarns. Fibers 6, 2 (2018), 35.Google ScholarGoogle ScholarCross RefCross Ref
  57. Dorothy Anne Hardy, Zahra Rahemtulla, Achala Satharasinghe, Arash Shahidi, Carlos Oliveira, Ioannis Anastasopoulos, Mohamad Nour Nashed, Matholo Kgatuke, Abiodun Komolafe, Russel Torah, 2020. Wash testing of electronic yarn. Materials 13, 5 (2020), 1228.Google ScholarGoogle ScholarCross RefCross Ref
  58. Chie Hayakawa, Shinya Funakawa, Kazumichi Fujii, Atsunobu Kadono, and Takashi Kosaki. 2014. Effects of climatic and soil properties on cellulose decomposition rates in temperate and tropical forests. Biology and fertility of soils 50 (2014), 633–643.Google ScholarGoogle Scholar
  59. J Hayward. 2020. E-Textiles and Smart Clothing 2020–2030: Technologies. Markets and Players (2020).Google ScholarGoogle Scholar
  60. Julian Hildebrandt, Philipp Brauner, and Martina Ziefle. 2015. Smart textiles as intuitive and ubiquitous user interfaces for smart homes. In Human Aspects of IT for the Aged Population. Design for Everyday Life: First International Conference, ITAP 2015, Held as Part of HCI International 2015, Los Angeles, CA, USA, August 2-7, 2015. Proceedings, Part II 1. Springer, 423–434.Google ScholarGoogle ScholarCross RefCross Ref
  61. Md Milon Hossain, Braden M Li, Busra Sennik, Jesse S Jur, and Philip D Bradford. 2022. Adhesive free, conformable and washable carbon nanotube fabric electrodes for biosensing. npj Flexible Electronics 6, 1 (2022), 97.Google ScholarGoogle Scholar
  62. Elaine Igoe. 2021. Textile Design Theory in the Making. Bloomsbury Publishing.Google ScholarGoogle Scholar
  63. Bio inspired Textiles. 2023. Bio-inspired Textiles. Retrieved September, 2023 from https://www.bioinspiredtextiles.com/Google ScholarGoogle Scholar
  64. Ebru Kurbak Irene Posh. 2018. The embroidered computer. Retrieved September, 2023 from https://ireneposch.net/the-embroidered-computer/Google ScholarGoogle Scholar
  65. E Jaafar, Muhammad Kashif, Siti Kudnie Sahari, and Zainab Ngaini. 2018. Study on morphological, optical and electrical properties of graphene oxide (GO) and reduced graphene oxide (rGO). In Materials Science Forum, Vol. 917. Trans Tech Publ, 112–116.Google ScholarGoogle Scholar
  66. Aayush Kumar Jaiswal, Vinay Kumar, Elina Jansson, Olli-Heikki Huttunen, Akio Yamamoto, Minna Vikman, Alexey Khakalo, Jussi Hiltunen, and Mohammad H Behfar. 2023. Biodegradable Cellulose Nanocomposite Substrate for Recyclable Flexible Printed Electronics. Advanced Electronic Materials 9, 4 (2023), 2201094.Google ScholarGoogle ScholarCross RefCross Ref
  67. Ranjith Jayasekara, Ian Harding, Ian Bowater, Gregor BY Christie, and Greg T Lonergan. 2003. Biodegradation by composting of surface modified starch and PVA blended films. Journal of Polymers and the Environment 11 (2003), 49–56.Google ScholarGoogle ScholarCross RefCross Ref
  68. Yimin Ji, Yuzhou Li, Guoqiang Chen, and Tieling Xing. 2017. Fire-resistant and highly electrically conductive silk fabrics fabricated with reduced graphene oxide via dry-coating. Materials & Design 133 (2017), 528–535.Google ScholarGoogle ScholarCross RefCross Ref
  69. ML Juul Søndergaard and M Balaam. 2022. Editorial: Designing with Bodily Materials. Proceedings of DRS (2022).Google ScholarGoogle Scholar
  70. Pirjo Kääriäinen, Liisa Tervinen, Tapani Vuorinen, Nina Riutta, 2020. The CHEMARTS cookbook. Aalto University.Google ScholarGoogle Scholar
  71. Naba Kumar Kalita, Ninad Anil Damare, Doli Hazarika, Purabi Bhagabati, Ajay Kalamdhad, and Vimal Katiyar. 2021. Biodegradation and characterization study of compostable PLA bioplastic containing algae biomass as potential degradation accelerator. Environmental Challenges 3 (2021), 100067.Google ScholarGoogle ScholarCross RefCross Ref
  72. Armağan Karahanoğlu and Çiğdem Erbuğ. 2011. Perceived qualities of smart wearables: determinants of user acceptance. In Proceedings of the 2011 conference on designing pleasurable products and interfaces. 1–8.Google ScholarGoogle Scholar
  73. Elvin Karana, Bahareh Barati, Valentina Rognoli, Anouk Zeeuw Van Der Laan, 2015. Material driven design (MDD): A method to design for material experiences. International journal of design 9, 2 (2015), 35–54.Google ScholarGoogle Scholar
  74. Ingrid M Keseler, César Bonavides-Martínez, Julio Collado-Vides, Socorro Gama-Castro, Robert P Gunsalus, D Aaron Johnson, Markus Krummenacker, Laura M Nolan, Suzanne Paley, Ian T Paulsen, 2009. EcoCyc: a comprehensive view of Escherichia coli biology. Nucleic acids research 37, suppl_1 (2009), D464–D470.Google ScholarGoogle Scholar
  75. Simon Knoch, Francine Pelletier, Mikaël Larose, Gérald Chouinard, Marie-Josée Dumont, and Jason R Tavares. 2020. Surface modification of PLA nets intended for agricultural applications. Colloids and Surfaces A: Physicochemical and Engineering Aspects 598 (2020), 124787.Google ScholarGoogle ScholarCross RefCross Ref
  76. Andreas R Köhler. 2013. Challenges for eco-design of emerging technologies: The case of electronic textiles. Materials & Design 51 (2013), 51–60.Google ScholarGoogle ScholarCross RefCross Ref
  77. Majeti NV Ravi Kumar. 2000. A review of chitin and chitosan applications. Reactive and functional polymers 46, 1 (2000), 1–27.Google ScholarGoogle Scholar
  78. Stacey Kuznetsov, Cassandra Barrett, Piyum Fernando, and Kat Fowler. 2018. Antibiotic-responsive bioart: Exploring DIYbio as a design studio practice. In Proceedings of the 2018 CHI Conference on Human Factors in Computing Systems. 1–14.Google ScholarGoogle ScholarDigital LibraryDigital Library
  79. Healthy Materials Lab. 2024.. Retrieved January, 2024 from https://healthymaterialslab.org/Google ScholarGoogle Scholar
  80. E Lazaro and K Vega. 2019. From Plastic to Biomaterials: Prototyping DIY Electronics with Mycelium. In Proceedings of the UbiComp/ISWC ‘19 Adjunct: Adjunct Proceedings of the 2019 ACM International Joint Conference on Pervasive and Ubiquitous Computing and Proceedings of the 2019 ACM International Symposium on Wearable Computers. ACM London, UK.Google ScholarGoogle Scholar
  81. Eldy S Lazaro Vasquez, Lily M Gabriel, Mikhaila Friske, Shanel Wu, Sasha De Koninck, Laura Devendorf, and Mirela Alistar. 2023. Designing Dissolving Wearables. In Adjunct Proceedings of the 2023 ACM International Joint Conference on Pervasive and Ubiquitous Computing & the 2023 ACM International Symposium on Wearable Computing. 286–290.Google ScholarGoogle Scholar
  82. Eldy S Lazaro Vasquez, Netta Ofer, Shanel Wu, Mary Etta West, Mirela Alistar, and Laura Devendorf. 2022. Exploring Biofoam as a Material for Tangible Interaction. In Designing Interactive Systems Conference. 1525–1539.Google ScholarGoogle ScholarDigital LibraryDigital Library
  83. Eldy S Lazaro Vasquez, Hao-Chuan Wang, and Katia Vega. 2020. Introducing the sustainable prototyping life cycle for digital fabrication to designers. In Proceedings of the 2020 ACM Designing Interactive Systems Conference. 1301–1312.Google ScholarGoogle ScholarDigital LibraryDigital Library
  84. Suzanne Lee, Amy Congdon, Georgia Parker, and Charlotte Borst. 2020. Understanding "Bio" Material Innovations: A primer for the fashion industry. Technical Report.Google ScholarGoogle Scholar
  85. Zhouyue Lei and Peiyi Wu. 2021. Bioinspired quasi-solid ionic conductors: materials, processing, and applications. Accounts of Materials Research 2, 12 (2021), 1203–1214.Google ScholarGoogle ScholarCross RefCross Ref
  86. Steven K Leist, Dajing Gao, Richard Chiou, and Jack Zhou. 2017. Investigating the shape memory properties of 4D printed polylactic acid (PLA) and the concept of 4D printing onto nylon fabrics for the creation of smart textiles. Virtual and Physical Prototyping 12, 4 (2017), 290–300.Google ScholarGoogle ScholarCross RefCross Ref
  87. Xiuhong Li, Shuang Chen, Yujie Peng, Zhong Zheng, Jing Li, and Fei Zhong. 2022. Materials, preparation strategies, and wearable sensor applications of conductive fibers: A review. Sensors 22, 8 (2022), 3028.Google ScholarGoogle ScholarCross RefCross Ref
  88. Jianwei Liang, Xiaoning Zhang, Chi Yan, Yixuan Wang, Michael L Norton, Xijun Wei, Carrie Donley, Yong Zhu, Peng Xiao, and Yunhuai Zhang. 2020. Preparation and enhanced supercapacitance performance of carbonized silk by feeding silkworms MoO2 nanoparticles. Materials & Design 196 (2020), 109137.Google ScholarGoogle ScholarCross RefCross Ref
  89. Xiaoping Liang, Mengjia Zhu, Haifang Li, Jinxin Dou, Muqiang Jian, Kailun Xia, Shuo Li, and Yingying Zhang. 2022. Hydrophilic, breathable, and washable graphene decorated textile assisted by silk sericin for integrated multimodal smart wearables. Advanced Functional Materials 32, 42 (2022), 2200162.Google ScholarGoogle ScholarCross RefCross Ref
  90. Yong Liu, Dingshan Yu, Chao Zeng, Zongcheng Miao, and Liming Dai. 2010. Biocompatible graphene oxide-based glucose biosensors. Langmuir 26, 9 (2010), 6158–6160.Google ScholarGoogle ScholarCross RefCross Ref
  91. Zhuomiao Liu, Jian Zhao, Kun Lu, Zhenyu Wang, Liyun Yin, Hao Zheng, Xiao Wang, Liang Mao, and Baoshan Xing. 2022. Biodegradation of graphene oxide by insects (tenebrio molitor larvae): role of the gut microbiome and enzymes. Environmental Science & Technology 56, 23 (2022), 16737–16747.Google ScholarGoogle ScholarCross RefCross Ref
  92. Jian Lv, Itthipon Jeerapan, Farshad Tehrani, Lu Yin, Cristian Abraham Silva-Lopez, Ji-Hyun Jang, Davina Joshuia, Rushabh Shah, Yuyan Liang, Lingye Xie, 2018. Sweat-based wearable energy harvesting-storage hybrid textile devices. Energy & Environmental Science 11, 12 (2018), 3431–3442.Google ScholarGoogle ScholarCross RefCross Ref
  93. E MacArthur. 2017. Foundation A New Textiles Economy: Redesigning Fashion’s Future. London, UK (2017).Google ScholarGoogle Scholar
  94. Josefa Fabiana Martucci and Roxana Alejandra Ruseckaite. 2009. Biodegradation of three-layer laminate films based on gelatin under indoor soil conditions. Polymer Degradation and Stability 94, 8 (2009), 1307–1313.Google ScholarGoogle ScholarCross RefCross Ref
  95. Materiom. 2024. Materiom. Retrieved January, 2024 from https://materiom.org/Google ScholarGoogle Scholar
  96. Sara Mlakar, Mira Alida Haberfellner, Hans-Christian Jetter, and Michael Haller. 2021. Exploring affordances of surface gestures on textile user interfaces. In Designing Interactive Systems Conference 2021. 1159–1170.Google ScholarGoogle ScholarDigital LibraryDigital Library
  97. Debashish Mohanta, Soma Patnaik, Sanchit Sood, and Nilanjan Das. 2019. Carbon nanotubes: Evaluation of toxicity at biointerfaces. Journal of pharmaceutical analysis 9, 5 (2019), 293–300.Google ScholarGoogle ScholarCross RefCross Ref
  98. Manideepa Mukherjee. 2019. Challenges and opportunities of textile based smart sanitary napkin design. In Adjunct Proceedings of the 2019 ACM International Joint Conference on Pervasive and Ubiquitous Computing and Proceedings of the 2019 ACM International Symposium on Wearable Computers. 1044–1046.Google ScholarGoogle ScholarDigital LibraryDigital Library
  99. Tae Jin Mun, Shi Hyeong Kim, Jong Woo Park, Ji Hwan Moon, Yongwoo Jang, Chi Huynh, Ray H Baughman, and Seon Jeong Kim. 2020. Wearable energy generating and storing textile based on carbon nanotube yarns. Advanced Functional Materials 30, 23 (2020), 2000411.Google ScholarGoogle ScholarCross RefCross Ref
  100. MycoWorks. 2023. MycoWorks. Retrieved September, 2023 from https://www.mycoworks.comGoogle ScholarGoogle Scholar
  101. Nasa. 2023. Technology Readiness Level. Retrieved September, 2023 from https://www.nasa.gov/directorates/heo/scan/engineering/technology/technology_readiness_levelGoogle ScholarGoogle Scholar
  102. United Nations. 2023. THE 17 GOALS. Retrieved September, 2023 from https://sdgs.un.org/goalsGoogle ScholarGoogle Scholar
  103. Audrey Ng. 2017. Grown microbial 3D fiber art, Ava: fusion of traditional art with technology. In Proceedings of the 2017 ACM International Symposium on Wearable Computers. 209–214.Google ScholarGoogle ScholarDigital LibraryDigital Library
  104. Kirsi Niinimäki. 2017. Fashion in a circular economy. Springer.Google ScholarGoogle Scholar
  105. Delphine Notta-Cuvier, Jérémy Odent, Rémi Delille, Marius Murariu, Franck Lauro, Jean-Marie Raquez, Bruno Bennani, and Philippe Dubois. 2014. Tailoring polylactide (PLA) properties for automotive applications: Effect of addition of designed additives on main mechanical properties. Polymer Testing 36 (2014), 1–9.Google ScholarGoogle ScholarCross RefCross Ref
  106. Netta Ofer and Mirela Alistar. 2023. Felt Experiences with Kombucha Scoby: Exploring First-person Perspectives with Living Matter. In Proceedings of the 2023 CHI Conference on Human Factors in Computing Systems. 1–18.Google ScholarGoogle ScholarDigital LibraryDigital Library
  107. Netta Ofer, Fiona Bell, and Mirela Alistar. 2021. Designing direct interactions with bioluminescent algae. In Designing Interactive Systems Conference 2021. 1230–1241.Google ScholarGoogle Scholar
  108. SHW Ossevoort. 2013. Improving the sustainability of smart textiles. In Multidisciplinary know-how for smart-textiles developers. Elsevier, 399–419.Google ScholarGoogle Scholar
  109. Yash Hemant Pandya, Manish Bakshi, Anushka Sharma, H Pandya, and H Pandya. 2022. Agar-Agar extraction, structural properties and applications: A review. Pharma Innov J 11 (2022), 1151–1157.Google ScholarGoogle Scholar
  110. Shomini Parashar and Narendra Kumar. 2018. Studies on agarolytic bacterial isolates from agricultural and industrial soil. Iranian Journal of Microbiology 10, 5 (2018), 324.Google ScholarGoogle Scholar
  111. Sungjin Park, Nihar Mohanty, Ji Won Suk, Ashvin Nagaraja, Jinho An, Richard D Piner, Weiwei Cai, Daniel R Dreyer, Vikas Berry, and Rodney S Ruoff. 2010. Biocompatible, robust free-standing paper composed of a TWEEN/graphene composite. Advanced Materials 22, 15 (2010), 1736–1740.Google ScholarGoogle ScholarCross RefCross Ref
  112. European Parlament. 2023. The impact of textile production and waste on the environment (infographics). Retrieved September, 2023 from https://www.europarl.europa.eu/news/en/headlines/society/20201208STO93327/the-impact-of-textile-production-and-waste-on-the-environment-infographicsGoogle ScholarGoogle Scholar
  113. Gabriele Pizzino, Natasha Irrera, Mariapaola Cucinotta, Giovanni Pallio, Federica Mannino, Vincenzo Arcoraci, Francesco Squadrito, Domenica Altavilla, Alessandra Bitto, 2017. Oxidative stress: harms and benefits for human health. Oxidative medicine and cellular longevity 2017 (2017).Google ScholarGoogle Scholar
  114. Emmi Pouta 2023. Layered Approaches-Woven eTextile Explorations Through Applied Textile Thinking. (2023).Google ScholarGoogle Scholar
  115. Emmi Pouta and Jussi Ville Mikkonen. 2022. Woven eTextiles in HCI—a Literature Review. In Designing Interactive Systems Conference. 1099–1118.Google ScholarGoogle ScholarDigital LibraryDigital Library
  116. Emmi Pouta, Riia Vidgren, Jaana Vapaavuori, and Mithila Mohan. 2022. Intertwining material science and textile thinking: Aspects of contrast and collaboration. (2022).Google ScholarGoogle Scholar
  117. STARTS Programme. 2023. reFream. Retrieved September, 2023 from https://re-fream.eu/Google ScholarGoogle Scholar
  118. Jiesheng Ren, Chaoxia Wang, Xuan Zhang, Tian Carey, Kunlin Chen, Yunjie Yin, and Felice Torrisi. 2017. Environmentally-friendly conductive cotton fabric as flexible strain sensor based on hot press reduced graphene oxide. Carbon 111 (2017), 622–630.Google ScholarGoogle ScholarCross RefCross Ref
  119. Jan-Georg Rosenboom, Robert Langer, and Giovanni Traverso. 2022. Bioplastics for a circular economy. Nature Reviews Materials 7, 2 (2022), 117–137.Google ScholarGoogle ScholarCross RefCross Ref
  120. Sigrid Rotzler, Christine Kallmayer, Christian Dils, Malte von Krshiwoblozki, Ulrich Bauer, and Martin Schneider-Ramelow. 2020. Improving the washability of smart textiles: Influence of different washing conditions on textile integrated conductor tracks. The Journal of The Textile Institute 111, 12 (2020), 1766–1777.Google ScholarGoogle ScholarCross RefCross Ref
  121. Nurul Saadah Said and Norizah Mhd Sarbon. 2022. Physical and mechanical characteristics of gelatin-based films as a potential food packaging material: A review. Membranes 12, 5 (2022), 442.Google ScholarGoogle ScholarCross RefCross Ref
  122. Marco Santoro, Sarita R Shah, Jennifer L Walker, and Antonios G Mikos. 2016. Poly (lactic acid) nanofibrous scaffolds for tissue engineering. Advanced drug delivery reviews 107 (2016), 206–212.Google ScholarGoogle Scholar
  123. Abhilash Sasidharan, LS Panchakarla, Parwathy Chandran, Deepthy Menon, Shantikumar Nair, CNR Rao, and Manzoor Koyakutty. 2011. Differential nano-bio interactions and toxicity effects of pristine versus functionalized graphene. Nanoscale 3, 6 (2011), 2461–2464.Google ScholarGoogle ScholarCross RefCross Ref
  124. Karsten Schischke, Nils F Nissen, and Martin Schneider-Ramelow. 2020. Flexible, stretchable, conformal electronics, and smart textiles: Environmental life cycle considerations for emerging applications. MRS Communications 10, 1 (2020), 69–82.Google ScholarGoogle ScholarCross RefCross Ref
  125. HaoTian Harvey Shi, Yifei Pan, Lin Xu, Xueming Feng, Wenyu Wang, Prasad Potluri, Liangbing Hu, Tawfique Hasan, and Yan Yan Shery Huang. 2023. Sustainable electronic textiles towards scalable commercialization. Nature Materials (2023), 1–10.Google ScholarGoogle Scholar
  126. Softwear. 2023. Softwear-dn. Retrieved September, 2023 from https://softwear-dn.eu/Google ScholarGoogle Scholar
  127. Katherine Wei Song and Eric Paulos. 2023. Vim: Customizable, Decomposable Electrical Energy Storage. In Proceedings of the 2023 CHI Conference on Human Factors in Computing Systems. 1–18.Google ScholarGoogle ScholarDigital LibraryDigital Library
  128. Giorgio Speranza. 2019. The role of functionalization in the applications of carbon materials: An overview. C 5, 4 (2019), 84.Google ScholarGoogle ScholarCross RefCross Ref
  129. Patritsia Stathatou, Alysia Garmulewicz, Liz Corbin, Pilar Bolumburu, and Zoe Kremer. 2022. Biomaterials and Regenerative Agriculture: Linkages and Opportunities: The Case of the Great Lakes Region, Michigan. Technical Report.Google ScholarGoogle Scholar
  130. Patritsia Maria Stathatou, Liz Corbin, J Carson Meredith, and Alysia Garmulewicz. 2023. Biomaterials and Regenerative Agriculture: A Methodological Framework to Enable Circular Transitions. Sustainability 15, 19 (2023), 14306.Google ScholarGoogle ScholarCross RefCross Ref
  131. Sharifjon Sulaymonov and Sh A Kholboeva. 2023. OEKO-TEX® STANDARD 100 TEXTILE PRODUCT SAFETY MANAGEMENT SYSTEM ROLE IN PRODUCT QUALITY ASSESSMENT ACCORDING TO REQUIREMENTS. International Bulletin of Applied Science and Technology 3, 5 (2023), 352–360.Google ScholarGoogle Scholar
  132. Qingqing Sun, Binbin Qian, Koichiro Uto, Jinzhou Chen, Xuying Liu, and Takeo Minari. 2018. Functional biomaterials towards flexible electronics and sensors. Biosensors and Bioelectronics 119 (2018), 237–251.Google ScholarGoogle ScholarCross RefCross Ref
  133. Mathias Sundholm, Jingyuan Cheng, Bo Zhou, Akash Sethi, and Paul Lukowicz. 2014. Smart-mat: Recognizing and counting gym exercises with low-cost resistive pressure sensing matrix. In Proceedings of the 2014 ACM international joint conference on pervasive and ubiquitous computing. 373–382.Google ScholarGoogle ScholarDigital LibraryDigital Library
  134. Yasuhiko Tabata. 2009. Biomaterial technology for tissue engineering applications. Journal of the Royal Society interface 6, suppl_3 (2009), S311–S324.Google ScholarGoogle Scholar
  135. Riikka Talman. 2019. Changeability as a quality in textile design. Ph. D. Dissertation. Högskolan i Borås.Google ScholarGoogle Scholar
  136. Mahmut Tas, Yasin Altin, and Ayse Bedeloglu. 2019. Graphene and graphene oxide-coated polyamide monofilament yarns for fiber-shaped flexible electrodes. The journal of the Textile Institute 110, 1 (2019), 67–73.Google ScholarGoogle ScholarCross RefCross Ref
  137. Xiao Tian and Tao Hua. 2021. Antibacterial, scalable manufacturing, skin-attachable, and eco-friendly fabric triboelectric nanogenerators for self-powered sensing. ACS Sustainable Chemistry & Engineering 9, 39 (2021), 13356–13366.Google ScholarGoogle ScholarCross RefCross Ref
  138. Granch Berhe Tseghai, Desalegn Alemu Mengistie, Benny Malengier, Kinde Anlay Fante, and Lieva Van Langenhove. 2020. PEDOT:PSS-Based Conductive Textiles and Their Applications. Sensors 20, 7 (2020). https://doi.org/10.3390/s20071881Google ScholarGoogle ScholarCross RefCross Ref
  139. Deniz Tümerdem and Leman Figen Gül. 2023. Temporalities of a DIY biocomposite through material exploration. The Design Journal (2023), 1–19.Google ScholarGoogle Scholar
  140. Aalto University. 2023. Bioinnovation Center. Retrieved September, 2023 from https://www.aalto.fi/en/aalto-university-bioinnovation-center?check_logged_in=1Google ScholarGoogle Scholar
  141. Leonardo Hidalgo Uribe. 2023. Making colour a relational practice: correspondence between materials and local environments in natural textile dyeing processes. In Biocolours: Sustainable stories from nature, lab and industry. Aalto University, 107–133.Google ScholarGoogle Scholar
  142. Natascha M Van der Velden, Kristi Kuusk, and Andreas R Köhler. 2015. Life cycle assessment and eco-design of smart textiles: The importance of material selection demonstrated through e-textile product redesign. Materials & Design 84 (2015), 313–324.Google ScholarGoogle ScholarCross RefCross Ref
  143. Eldy S Lazaro Vasquez and Katia Vega. 2019. Myco-accessories: sustainable wearables with biodegradable materials. In Proceedings of the 2019 ACM International Symposium on Wearable Computers. 306–311.Google ScholarGoogle ScholarDigital LibraryDigital Library
  144. Paula Veske and Elina Ilén. 2021. Review of the end-of-life solutions in electronics-based smart textiles. The Journal of the Textile Institute 112, 9 (2021), 1500–1513.Google ScholarGoogle ScholarCross RefCross Ref
  145. Paula Veske, Kristi Kuusk, Marina Toeters, and Barbro Scholz. 2019. Environmental sustainability of e-textile products approached by makers and manufacturers. (2019).Google ScholarGoogle Scholar
  146. Mahmoud Wagih, Sheng Yong, Kai Yang, Alex S Weddell, and Steve Beeby. 2022. Printed non-metallic textile-based carbon antenna for low-cost green wearable applications. In 2022 16th European Conference on Antennas and Propagation (EuCAP). IEEE, 1–4.Google ScholarGoogle ScholarCross RefCross Ref
  147. Ruoxing Wang, Jiajie Sui, and Xudong Wang. 2022. Natural piezoelectric biomaterials: a biocompatible and sustainable building block for biomedical devices. ACS nano 16, 11 (2022), 17708–17728.Google ScholarGoogle Scholar
  148. Wen Wang, Lining Yao, Chin-Yi Cheng, Teng Zhang, Hiroshi Atsumi, Luda Wang, Guanyun Wang, Oksana Anilionyte, Helene Steiner, Jifei Ou, 2017. Harnessing the hygroscopic and biofluorescent behaviors of genetically tractable microbial cells to design biohybrid wearables. Science advances 3, 5 (2017), e1601984.Google ScholarGoogle Scholar
  149. Zhiting Wei, Jinlong Wang, Yanhua Liu, Jinxia Yuan, Tao Liu, Guoli Du, Siqiyuan Zhu, and Shuangxi Nie. 2022. Sustainable triboelectric materials for smart active sensing systems. Advanced Functional Materials 32, 52 (2022), 2208277.Google ScholarGoogle ScholarCross RefCross Ref
  150. Alejandro Javier Weiss Münchmeyer and María José Besoain Narvaez. 2022. Biomateriales basados en el territorio: Metodología para la creación de una paleta biomaterial situada/Territory-based biomaterials: Methodology to create a situated biomaterial palette. (2022).Google ScholarGoogle Scholar
  151. Robin Whittemore and Kathleen Knafl. 2005. The integrative review: updated methodology. Journal of advanced nursing 52, 5 (2005), 546–553.Google ScholarGoogle ScholarCross RefCross Ref
  152. Rachael Wickenden. 2021. Rethinking E-textile design: process, purpose and sustainability. Nottingham Trent University (United Kingdom).Google ScholarGoogle Scholar
  153. David Franklyn Williams and David Franklyn Williams. 1999. The Williams dictionary of biomaterials. Liverpool University Press.Google ScholarGoogle Scholar
  154. Claes Wohlin. 2014. Guidelines for snowballing in systematic literature studies and a replication in software engineering. In Proceedings of the 18th international conference on evaluation and assessment in software engineering. 1–10.Google ScholarGoogle ScholarDigital LibraryDigital Library
  155. Natalia Wrońska, Nadia Katir, Marta Nowak-Lange, Abdelkrim El Kadib, and Katarzyna Lisowska. 2023. Biodegradable Chitosan-Based Films as an Alternative to Plastic Packaging. Foods 12, 18 (2023), 3519.Google ScholarGoogle ScholarCross RefCross Ref
  156. Shanel Wu and Laura Devendorf. 2020. Unfabricate: designing smart textiles for disassembly. In proceedings of the 2020 CHI conference on human factors in computing systems. 1–14.Google ScholarGoogle ScholarDigital LibraryDigital Library
  157. Wenzheng Wu, Wenli Ye, Zichao Wu, Peng Geng, Yulei Wang, and Ji Zhao. 2017. Influence of layer thickness, raster angle, deformation temperature and recovery temperature on the shape-memory effect of 3D-printed polylactic acid samples. Materials 10, 8 (2017), 970.Google ScholarGoogle ScholarCross RefCross Ref
  158. Geng Yang, Jia Deng, Gaoyang Pang, Hao Zhang, Jiayi Li, Bin Deng, Zhibo Pang, Juan Xu, Mingzhe Jiang, Pasi Liljeberg, 2018. An IoT-enabled stroke rehabilitation system based on smart wearable armband and machine learning. IEEE journal of translational engineering in health and medicine 6 (2018), 1–10.Google ScholarGoogle ScholarCross RefCross Ref
  159. Mei Yang and Minfang Zhang. 2019. Biodegradation of carbon nanotubes by macrophages. Frontiers in Materials 6 (2019), 225.Google ScholarGoogle ScholarCross RefCross Ref
  160. Yadie Yang, Minglonghai Zhang, Zixin Ju, Po Ying Tam, Tao Hua, Muhammad Waseem Younas, Hasan Kamrul, and Hong Hu. 2021. Poly (lactic acid) fibers, yarns and fabrics: Manufacturing, properties and applications. Textile Research Journal 91, 13-14 (2021), 1641–1669.Google ScholarGoogle ScholarCross RefCross Ref
  161. Lining Yao, Jifei Ou, Chin-Yi Cheng, Helene Steiner, Wen Wang, Guanyun Wang, and Hiroshi Ishii. 2015. BioLogic: natto cells as nanoactuators for shape changing interfaces. In Proceedings of the 33rd Annual ACM Conference on Human Factors in Computing Systems. 1–10.Google ScholarGoogle ScholarDigital LibraryDigital Library
  162. Xiaopei Zhang, Amal Al-Dossary, Myer Hussain, Peter Setlow, and Jiahe Li. 2020. Applications of Bacillus subtilis spores in biotechnology and advanced materials. Applied and Environmental Microbiology 86, 17 (2020), e01096–20.Google ScholarGoogle ScholarCross RefCross Ref

Index Terms

  1. Towards More Sustainable Interactive Textiles: A Literature Review on The Use of Biomaterials for eTextiles.

    Recommendations

    Comments

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Sign in
    • Article Metrics

      • Downloads (Last 12 months)358
      • Downloads (Last 6 weeks)358

      Other Metrics

    PDF Format

    View or Download as a PDF file.

    PDF

    eReader

    View online with eReader.

    eReader

    HTML Format

    View this article in HTML Format .

    View HTML Format