• Ei tuloksia

Casein-magnesium composite as an intumescent fire retardant coating for wood

N/A
N/A
Info
Lataa
Protected

Academic year: 2022

Jaa "Casein-magnesium composite as an intumescent fire retardant coating for wood"

Copied!
25
0
0

Kokoteksti

(1)

UEF//eRepository

DSpace https://erepo.uef.fi

Rinnakkaistallenteet Luonnontieteiden ja metsätieteiden tiedekunta

2020

Casein-magnesium composite as an

intumescent fire retardant coating for wood

Uddin, Mezbah

Elsevier BV

Tieteelliset aikakauslehtiartikkelit

© Elsevier Ltd.

CC BY-NC-ND https://creativecommons.org/licenses/by-nc-nd/4.0/

http://dx.doi.org/10.1016/j.firesaf.2019.102943

https://erepo.uef.fi/handle/123456789/7983

Downloaded from University of Eastern Finland's eRepository

(2)

Journal Pre-proof

Casein-magnesium composite as an intumescent fire retardant coating for wood Mezbah Uddin, Kalle Kiviranta, Sari Suvanto, Leila Alvila, Jari Leskinen, Reijo Lappalainen, Antti Haapala

PII: S0379-7112(19)30135-3

DOI: https://doi.org/10.1016/j.firesaf.2019.102943 Reference: FISJ 102943

To appear in: Fire Safety Journal Received Date: 17 March 2019 Revised Date: 20 December 2019 Accepted Date: 27 December 2019

Please cite this article as: M. Uddin, K. Kiviranta, S. Suvanto, L. Alvila, J. Leskinen, R. Lappalainen, A.

Haapala, Casein-magnesium composite as an intumescent fire retardant coating for wood, Fire Safety Journal (2020), doi: https://doi.org/10.1016/j.firesaf.2019.102943.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.

(3)

1 1

Casein-magnesium composite as an intumescent fire retardant coating for wood 2

Mezbah Uddina, Kalle Kivirantab, Sari Suvantoc, Leila Alvilac, Jari Leskinend, Reijo 3

Lappalainend,e, Antti Haapalaa*

4

a School of Forest Sciences, University of Eastern Finland, FIN-80100, Joensuu, Finland 5

b Savonia University of Applied Sciences, FIN-70101, Kuopio, Finland 6

c Department of Chemistry, University of Eastern Finland, FIN-80100, Joensuu, Finland 7

d SIB Labs, University of Eastern Finland, Kuopio Campus, FIN-70211, Kuopio, Finland 8

e Department of Applied Physics, University of Eastern Finland, FIN-70211, Kuopio, Finland 9

Corresponding author: antti.haapala@uef.fi 10

11

ABSTRACT 12

Fire hazard associated to wood and wood-based products is a clear shortcoming of natural 13

building materials and hence the use of fire retardants is mandated by national building codes.

14

In this study, we have synthesized a simple and green fire retardant composite using 15

magnesium hydroxide (Mg(OH)2) with casein protein and tested its performance as self- 16

adhesive coating paste. The thermogravimetric analyses indicated that casein-magnesium 17

composites are able to produce thermally stable insulating magnesium oxide (MgO) that 18

suppresses the fire spread as a result of the synergistic combination of physical and chemical 19

actions—the intumescent action of casein, formation of char and thermally stable MgO, and 20

the simultaneous release of water vapor retained by the casein. The most effective 21

performance was obtained with a sample (coating weight of 1080 g/m2) for which the ignition 22

(4)

2 time of pinewood substrate was increased by 147%, while the smoke production rate and the 23

peak heat release rate were decreased by 53% and 30%, respectively.

24 25

Keywords 26

Casein; Flame retardancy; Magnesium hydroxide; Thermal analysis; Wood 27

Declaration of interest: none 28

29

1. Introduction 30

Wood is naturally a fire sensitive material because of its thermally active carbohydrate 31

polymeric constituents. Commercial fire retardants are abundantly available for wood products 32

as paints and surface coatings for e.g. in places where rigid gypsum boards and other 33

structural protections are not feasible but issues on performance, environmental sustainability 34

and cost remain to be improved for many contemporary solutions. For instance, boron-based 35

products have reportedly suffer from leaching problems [1], and the halogenated fire 36

retardants produce halogen gases known to potentially cause respiratory issues [1-5].

37

Similarly, polyurethane-based fire retardants release hydrogen cyanide and carbon monoxide, 38

which are toxic [6].

39

Several studies have been on modified wood materials [7,8], wood-plastic composite 40

materials [9,10] and halogen-free fire retardant compounds as coatings and nano-coating 41

agents. These include clay nanopapers [11] composites of waterborne polyurethane (WPU), 42

aluminum trihydroxide (ATH) and mica [12], TiO2/ZnO coatings [13] and exfoliated vermiculite- 43

sodium silicate composites [14]. Other approaches are also used to insert fire retardant 44

chemicals into wood. For example, the polymerization of aniline in wood veneers [15] and 45

(5)

3 impregnation using a mixture of ammonium polyphosphate, magnesium hydroxide and 46

aluminum hydroxide [16].Recently, Merk et al. [17,18] investigated the mineralization of wood 47

by calcium carbonate using different approaches for improving fire retardancy and Xiao et al.

48

[19] reported reduction of both heat release and smoke production using modified silica.

49

Similarly the uses of graphene [24], nanoclay [25] and synthetic sulfonamides [20] have been 50

reported. To reduce the flammability behavior of wood with non-toxic additives is a potentially 51

valuable research topic for the wood-converting industry [21] while similar development is 52

seen on thermoset resins [22,23] and other materials [24-28].

53

The most ecofriendly nitrogen- and phosphorous-containing fire retardants (FRs) are the 54

casein [29] and whey proteins [30], deoxyribonucleic acid (DNA) [32] and DNA with chitosan 55

[33]. Caseins have traditional applications outside the food sector as e.g. binding materials as 56

crosslinking agents for tissue engineering [29]. Casein belongs to a family of phosphoproteins 57

containing 0.7-0.9% of phosphorous, which exist with protein through a serine ester linkage 58

called phosphoserine. The solubility of these calcium phosphor-caseinate clusters depends on 59

pH and casein micelles typically form stable colloidal dispersions with very few disulfide bonds 60

[31]. As caseins consist of phosphorous they can produce phosphoric acid to dehydrate 61

polymers. This is applied in the fire retardancy of textiles [29,32], where casein is shown to 62

promote char formation through dehydration of cellulose, however, to obtain significant 63

performance, rather high concentrations in relation to the mass of protected substrate are 64

required. They also generate nitrogen-phosphorus synergy in the combustion process [33,34].

65

Magnesium hydroxide is used as a fire retardant, which is attributed to the endothermic 66

decomposition into magnesium oxide and water that causes ignition to delay because energy 67

from heat source is absorbed by the reaction. Concurrently the release of water vapor dilutes 68

(6)

4 the flammable gases that are generated and the produced magnesium oxide can function as a 69

thermal barrier [35].

70

The study focuses on to develop approaches to synthesize a novel and cost-effective fire 71

retardant (FR) composite by preparing casein protein suspension in the presence of sodium 72

hydroxide (NaOH) followed by reacting with magnesium hydroxide (Mg(OH)2). The composite 73

provides multiple modes of actions to protect fire spread with relatively low content of casein.

74

The Scots pine sapwood was coated using prepared composite paste from 700 to 1080 g/m2. 75

The penetration depth of the coating components onto wood and formation of chemical bonds 76

between coating composite constituents was also investigated. The thermal stability of the 77

composite was investigated using thermogravimetry (TG) and differential thermogravimetry 78

(DTG). Cone calorimetry and TG were used to assess the self-extinguishing property and 79

thermal stability of the coated wood samples.

80

2. Experiments 81

2.1. Materials 82

Scots pine (Pinus sylvestris L.) sapwood samples (100 × 100 × 20 mm) were conditioned at 83

(22°C and 50% relative humidity RH) for a minimum of two weeks before cutting and coating.

84

Sodium hydroxide (NaOH) (≥ 97% pellet; Merck, Burlington, USA), casein from bovine milk 85

(technical grade), and Mg(OH)2 (reagent grade, 95%), both from Merck KGaA (Darmstadt, 86

Germany), were used directly without further purification. Deionized water was used during the 87

experiment when needed.

88 89

2.2. Sample preparation 90

First, 50 grams of water in a 250 ml beaker was heated to 60°C and 10 g of casein 91

introduced, while stirring continuously. Then 0.4 g of NaOH was added while stirring for 10 min 92

(7)

5 to facilitate even mixing and formation of sodium caseinate salt to complete the base mix.

93

Different composite formulations were prepared by adding 4 to 16 g of Mg(OH)2 (see Table 1) 94

into the suspension following continuous mixing for 10 min at 70°C to form a composite paste.

95

In preparing the coatings for 700, 775, 990 and 1080 g/m2 coating basis weights doses of 4, 8, 96

12, and 16 grams of Mg(OH)2 were used. The reaction mixture was stirred for 5 min to cool, 97

and then the viscous composites were cast onto two wood pieces, each measuring 98

100×100×20 mm and climatized at 22°C and 50% relative humidity (RH). The composites 99

were allowed to cover the specimen surfaces and the coating layer was made uniform with a 100

help of an adjustable coating blade (Fig. 1). Coated samples were dried at room temperature 101

for 48 h, and then kept at 22°C and 50% RH in a fixed humidity chamber for 7 days.

102

103

Fig 1. Reference (left) and a sample of coated Scots pine sapwood.

104 105

The weight of coating applied onto the samples (g/m2) was calculated by subtracting the 106

weight of uncoated sample from the weight of the coated and dried sample. The composition 107

and applied weight of coatings onto wood (g/m2) are summarized in Table 1.

108 109

Table 1. Composition of the samples coating and their weight onto each sample 110

Sample # Composition of coating

composite Coating applied (g) Coating basis weight (g/m2)

(8)

6

Reference - - -

Sample 1 4 g Mg(OH)2, 0.4 g NaOH, 10 g

casein 7 700

Sample 2 8 g Mg(OH)2, 0.4 g NaOH, 10 g

casein 7.7 770

Sample 3 12 g Mg(OH)2, 0.4 g NaOH, 10

g casein 9.9 990

Sample 4 16 g Mg(OH)2, 0.4 g NaOH, 10

g casein 10.8 1080

111 112

2.3. Surface analysis 113

The surface of the prepared composites was investigated with a FTIR spectrometer Vertex 114

70 (Bruker, Leipzig, Germany) equipped with an ATR platinum diamond, a sensitive 2 × 2 mm 115

diamond crystal surface, and a sample detector RT-DLaTGS. The same ATR-FTIR parameters 116

recorded within the range 4000–650 cm-1 were used in the OPUS 6.5 software (Bruker, 117

Leipzig, Germany) throughout the measurements.

118 119

2.4. Microanalysis 120

The penetration capacity of the coating components and the morphology of the coated 121

sample were investigated using scanning electron microscopy-energy-dispersive X-ray 122

spectroscopy (SEM-EDS). A cross-section, including the borderline of the wood and coating, 123

was observed with a Zeiss SigmaHD|VP Field Emission SEM (Carl Zeiss NTS, Cambridge, 124

UK). Also, the Mg concentration was identified on both sides of the interface between the 125

coating and the wood substrate with EDS (Thermo NSS; Thermo Fisher Scientific, Madison, 126

WI, USA) line scan of 20 evenly distributed points along a 1.2 mm line. A small chip of samples 127

(9)

7 was attached to a standard 12 mm aluminum stub for SEM specimens using a piece of 128

double-sided copper adhesive tape and imaged without any sputter coating in low vacuum, N2 129

atmosphere at 30 Pa pressure using 10 kV acceleration voltage, a backscatter electron 130

detector, and a working distance of 8.3 mm.

131 132

2.5. Thermal gravimetric analysis 133

The thermal properties of the samples were studied using a thermogravimetric analyzer 134

(TGA/STDA 851e/LF/1100; Mettler Toledo, Greifensee, Switzerland). Approximately 10 mg of 135

dry coating were taken from the coated wood samples, along with a thin layer of wood 136

material, using a sharp blade to be used for TGA measurement. Thermogravimetric 137

measurement was run under a nitrogen flow rate of 50 ml/min, a heating rate of 10°C/min, and 138

the temperature range of 25 to 800°C. The sample (approx. 9–10 mg) in solid form was kept in 139

an open alumina pan (70 µl).

140 141

2.6. Cone calorimetric analysis 142

Flammability behavior of the samples was analyzed with a cone calorimeter (ConeTool 1;

143

SGS Govmark Ltd., Farmingdale, IL, USA) at a radiant flux of 50 kW/m2 based on the ISO 144

5660-2 Standard Test Method. The heat release rate was measured based on the oxygen 145

consumption and the flow rate in the combustion product stream. The samples were set up 146

horizontally in the sample holder, and only the surface of the samples was exposed.

147

148

3. Results and discussion 149

3.1. Surface properties of the composites 150

(10)

8 The following FTIR-ATR spectra (Fig. 2) belong to the surface of four different casein- 151

magnesium hydroxide composites. The peak at 3695 cm-1 corresponds to the stretching of OH 152

in the Mg(OH)2 [36].The broad bands at 3279 cm-1 are associated with the stretching of OH in 153

the protein structure. The band at 2920 cm-1 corresponds to the stretching of CH2 in the protein 154

skeleton. The large peak at 1632 cm-1 is attributed to the stretching of the C=O of the amide 155

bond, the so-called amide I group (alpha helix and beta sheet structure of proteins), while the 156

peak at 1531 cm-1 refers to the combination of the stretching of C-N and deformation bending 157

of N-H, the so-called amide II groups [37]. The signal at 1082 cm-1 corresponds to the 158

stretching of the C-O in the COO- groups. The band at 1238 cm-1 is attributed to the stretching 159

of the P=O in phosphodiesters based on the information from ref. [38] but it could be also 160

related to C-N bond.

161

162

Fig 2. FTIR-ATR absorbance spectra of casein and four prepared composites.

163 164

(11)

9 The intensity of all the assigned peaks decreased with the increasing concentration of 165

Mg(OH)2, the sole exception being the stretching of OH at 3695 cm-1. The reduction of 166

intensities indicates that magnesium may interact with the different active moieties of casein, 167

but only weakly as no chemical bond formation is observed. For instance, Mg may interact 168

with casein protein through the strong long-range interaction of PO2 in the phosphate of the 169

casein. In addition, Mg may interact with the carbonyl C=O and C-N bonds of casein. Along 170

with these, Mg may also interact with the basic region (amine) of the casein protein [39].

171

Another reason for the decreasing intensity of the peaks is that Mg(OH)2 covers the functional 172

moieties of the casein – however, the precise interfacial phenomena between the principal 173

components is considered beyond the scope of this study. Further study will focus on in the 174

direction of the precise interfacial phenomena between the principal components, as 175

mentioned here the focus was to illustrate the performance of the achieved composite.

176 177

3.2. Coating component penetration into wood 178

The following images show the morphology and elemental depth profiles of the coated wood 179

samples. In Fig. 3a, the arrangement of mineral coating particles on top of the wood substrate 180

is shown. The coating seems rather uniform, but the layer thickness has some variation, as it 181

was not optimized; rather, the content of solids in the coating chemical was controlled. The Mg 182

profile measured with EDS is highlighted in white. The Mg concentration of 15.9 ± 3.1% (w/w) 183

was measured within the coating (1080 g/m2 coating). The measured Mg concentration profile 184

revealed that the studied fire retardant had penetrated about 560 µm into the wood.

185

The Mg concentration in the wood substrate varied locally from 2 to 12%. However, the local 186

minimum concentrations were measured at the hollow points on the surface. Therefore, it is 187

obvious that those values were greatly underestimated (Fig. 3a). The EDS analysis pathway 188

(12)

10 from wood to coating over their interface is shown with a yellow arrow. In Fig. 3b, the 189

composition of identified elements in the wood and coating layers is shown. The proof of 190

MgOH penetration into the wood surface was obtained with EDS as the Mg concentration 191

dropped to 0.7% or less below 500 µm in the wood, which, in practice, can be considered as 192

retardant free wood.

193

194

Fig 3. a) SEM image from the coating interface shows the concentration gradient of Mg with 195

some penetration into the wood surface. The yellow arrow shows the pathway measured by 196

EDS analysis, along which b) the elemental distribution of C, O, and Mg were measured.

197 198

3.3. Thermal stability analysis 199

The thermal stability of the samples was investigated using thermogravimetry (TG) and 200

differential thermogravimetry (DTG), as shown in Fig. 4a and 4b. In the case of the untreated 201

sample (only wood), weight loss occurred in three steps as observed from the both TGA and 202

DTGA curves. Bound water and volatile organic compounds are lost in the first step;

203

hemicellulose and cellulose are decomposed into CO2, CO, CH4, CH3OH, and CH3COOH in 204

(13)

11 the second step, called the charring step; and, finally, lignin is decomposed and char residue is 205

oxidized in the third step, called the calcining step [40,41].

206

In the case of the treated samples, weight loss occurred in four steps. First, water and 207

organic volatiles are lost as in the untreated sample. In the second step, hemicellulose and 208

cellulose are decomposed; weight loss is lower than for the untreated reference, which could 209

be due to the formation of synergy, which may modify the pyrolysis process by interrupting the 210

formation of the main degradation product levoglucosan from cellulose that facilitates thermally 211

stable char formation. Casein begins to decompose at 270°C and may start to form phosphoric 212

acid that may promote char formation [29].

213

In the third step, Mg(OH)2 starts to decompose at just under 350°C and forms thermally 214

stable MgO that protects the wood surface from burning. From the DTGA, it is obvious that 215

samples containing more Mg(OH)2 lose more weight at 415°C; this indicates that the more 216

substantial MgO layer formation effectively blocks the heat transfer. In the fourth step, lignin 217

starts to decompose at 425°C; however, the mass loss of all the samples coated with 218

composites is lower than that of the untreated reference because of the char and MgO layer 219

formation.

220

221

(14)

12 Fig 4. a) TG and (b) DTG curves of untreated wood and samples treated with coating 222

composites.

223 224

3.4. Flame retardant performance 225

To assess the flame retardancy of the samples, we monitored time to ignition (TTI) in 226

seconds, peak heat release rate (HRR) in kW/m2, total heat released (THR) in MJ/m2, average 227

smoke production (ASP) in m2/s, and total mass loss (TML) in g. The cone calorimetric test 228

was performed three times for each sample and the average value was used to assess the fire 229

retardancy. The results are shown in the following Table 2.

230 231

Table 2. Cone calorimetry results of the reference and coated wood (TTI: Time to ignition, 232

Peak HHR: Peak heat release rate, THR: Total heat released, TML: Total mass loss, ASP:

233

Average smoke production) 234

Sample TTI (s) Peak HHR (kW/m²) THR (MJ/m²) TML (g) ASP (m2/s)

Ref.* 12.1 216.0 79.5 46.6 0.0047

700 g/m2 11.5 147.7 64.3 42.4 0.0035

770 g/m2 17.5 140.0 61.9 41.5 0.0031

990 g/m2 26.3 127.3 56.1 42.5 0.0035

1080 g/m2 30.4 119.0 53.3 40.8 0.0022

* The reference used was uncoated pinewood 235

236

3.4.1. Time to ignition 237

TTI of all the samples was higher than the reference except for sample containing 700 g/m2 238

coating in Table 2, probably to inherent variation in wood composition. Sample containing the 239

highest amount of composite (1080 g/m2) delayed fire spreading significantly, which could be 240

due to the release of more water vapor during combustion which cools the surface of the 241

(15)

13 substrate. During the combustion process Mg(OH)2 also forms thermally stable magnesium 242

oxide (MgO) on the coating layer. Coatings have an intumescent functionality due to the 243

presence of casein, which prevents wood from burning by retaining inflammable gasses 244

between the wood and coating, like in [42], where the components produced a synergistic 245

effect in extending the ignition time. The synergistic effect may be caused through the complex 246

formation between casein and magnesium hydroxide. Divalent magnesium ion (Mg2+) forms 247

octahedral complex during reaction with casein protein, in which four bonds form with four 248

ligands of casein such as either aspartic acid (E) or glutamic acid (D). Another bond forms with 249

carboxylate oxygen atom of either glutamic acid (E) or aspartic acid (D), which functions as 250

monodentate ligand. One further bond forms with water molecule shown in the following figure 251

[45-47].

252

253

Fig 5. The interaction of magnesium with casein protein.

254 255

Trey et al. [15] found no improvement of TTI when applying polyaniline in veneers. Pan et al.

256

[16] similarly failed to show improvement in TTI by adding Mg, Al, or Si derivatives to nitrogen- 257

phosphorus fire retardants. Merk et al. [17] also reported no change in TTI for CaCO3 258

mineralized beech and spruce wood. Chang et al. [12] found that a WPU coating on wood 259

(16)

14 panels benefited from ATH and mica as fillers. The addition of inorganics resulted in an 260

extended TTI from 34 sec with WPU to 75 sec with ATH (WPU/ATH in 10/100 g), and the best 261

sample, with 15 g ATH and 85 g mica mixed in 10 g WPU, did not ignite at all. Their coating 262

system thickness was, in comparison approximately four times that of Sample 4 presented 263

here, making direct comparison difficult. However, a thicker coating generally provides better 264

flame retardant performance.

265

In addition, the study with a layer-by-layer coating using vermiculite and sodium silicate 266

(double layer) on wood substrate observed that the TTI was increased by 494% [14]. Film with 267

clay nanopaper increased the TTI of wood to 4.5 minutes; however, the heat flux in this study 268

differed from that of our study, being only 35 kW/m2 [11]. Although the performance is difficult 269

to compare directly due to differing test setups the displayed casein-magnesium composites 270

can clearly prolong the durability of wood in the presence of fire.

271 272

3.4.2. Heat release rate 273

While burning samples with a cone calorimeter, the untreated wood sample produces a 274

mixture of combustible gases, releases substantial heat, and forms a large exothermic peak 275

with a heat release rate (HRR) of 216 kW/m2. However, three exothermic peaks are formed for 276

all the treated samples: The first peak formed might be due to the intumescent behavior of the 277

coatings. When the treated samples are burned, Mg(OH)2 decomposes and produces water 278

vapor and MgO. This results in released heat and forms the second exothermic peak with a 279

low peak HRR. In the third combustion step, the char layer is decomposed, resulting in tiny 280

cracks that facilitate the evaporation of volatiles and release heat (the third exothermic peak).

281

The results indicate that the HRR of all the samples treated with composites containing 282

Mg(OH)2 is lower than the reference sample (Fig. 6) and decreases with the increasing 283

(17)

15 concentration of Mg(OH)2. The reason for the reduction of the HRR of the samples treated with 284

composites is the formation of thermally stable MgO and release of water from Mg(OH)2 285

decomposition inside the protective swelling coating layer of casein. Thermally stable MgO 286

may block the heat transfer through the material surface and results in char formation [12].

287

MgO may also block the heat transfer via suppressing the spread of oxygen gas and 288

combustible volatile compounds [16].

289

290

Fig 6. HRRs of the uncoated wood and coated wood samples.

291 292

3.4.3. Total heat released 293

The total heat released (THR) refers to the total amount of heat released from the materials 294

during combustion. Results indicate that the THR of all the treated samples is lower than that 295

of the untreated sample and that the THR is decreased with increasing the concentration of 296

Mg(OH)2. This is due to the fact that the samples containing more Mg(OH)2 produced a 297

considerable amount of thermally stable magnesium oxide which may absorb the heat.

298

Furthermore, water released from the decomposition of Mg(OH)2 is absorbed during 299

(18)

16 combustion, resulting in the total heat reduction [43,44]. The THR of coating containing 1080 300

g/m2 of composite was decreased by 41%.

301 302

3.4.4. Total mass loss 303

The total mass loss (TML) prior ignition is related to the moisture content of the coatings and 304

the water content of the wood itself [14]. The pine wood as a coating substrate contained ca.

305

9.5% of water (following the conditioning at RH 50%) and the moisture content in the coating 306

composites in these conditions was 4.5% of water (w/w). In terms of coated wood specimen, 307

the difference of moisture content between the lowest and the highest basis weight coatings (7 308

and 10.8 grams in absolute values) was hence less than 0.2 grams, which is not considered to 309

contribute significantly to their perceived mass loss. The TML of all the samples containing 310

Mg(OH)2 is lower than that of the untreated samples, and the TML of sample with the most 311

coating was the lowest (Table 2), which may be due to the higher char and thermally stable 312

magnesium oxide (MgO) formation (Fig. 4). Merk et al. [17] observed no significant change in 313

the TML for the CaCO3 mineralized beech and spruce wood. Several reasons can explain the 314

inconsistency of mass loss and smoke production trends observed between 770 and 990 g/m2 315

coatings. For instance, the anisotropic structure and possible differences in the wood substrate 316

composition, or small defects in some of the coating surfaces of 990 g/m2 specimen may have 317

contributed, but which we were regardless unable to identify from burnt samples.

318

Images of the surface of the samples after testing with cone calorimeter are shown in Fig. 7.

319

The char formation on the surface of the coated samples is clearly visible, but differences 320

among the coated samples were not significant, as the variation of mass loss between 321

samples was also limited.

322

(19)

17 323

Fig. 7. Untreated pinewood as a reference and the composite coated and charred pinewood 324

samples after a 600-second cone calorimetry test using a heat flux of 50 kW/m2. 325

326

3.4.5. Smoke production 327

The ASP decreases with increasing the concentration of Mg(OH)2. All the samples treated 328

with composite release less smoke than the untreated sample. During the combustion 329

process, casein-magnesium composite produced MgO layers having a high surface area, 330

which facilitates the absorption of smoke and smoke producer gasses [16]. Among all the 331

samples, sample containing 1080 g/m2 of composite releases the lowest amount of smoke, 332

which is decreased by about 53% compared with the untreated wood. The synergistic effect of 333

magnesium hydroxide and casein may suppress the smoke production similar to the study of 334

the wood sample impregnated with the mixture of 25% nitrogen–phosphorous fire retardant 335

and 10% magnesium and aluminum hydroxides [16].

336

4. Conclusions 337

Casein functions as an adhesive onto the wood surface and provides active sites for 338

interaction with magnesium while a pure casein suspension in reasonable concentration does 339

not improve the fire retardancy of wood and Mg(OH)2 alone lacks adhesive capacity to wood.

340

The deposition of magnesium in the pores of the pinewood at a 560 µm depth was confirmed 341

by SEM-EDS. Intumescent activity of the coating arises from casein, while the decomposition 342

of magnesium hydroxide at 415°C into thermally stable magnesium oxide and release of water 343

(20)

18 vapor provide additional protection. The cone calorimeter results showed that all the prepared 344

composites function as fire retardants and that the best performance was obtained with a 345

sample containing 1080 g/m2 of composite, extending TTI from 12 to 30 sec, while also 346

decreasing peak heat release from 216 to 119 kW/m2 and ASP from 0.005 to 0.002 m2/s, 347

respectively. Along with the thermally stable magnesium oxide and water vapor present, the 348

synergy obtained from casein and magnesium hydroxide composite coating can play a role for 349

improving the fire retardancy of wood products.

350 351

Acknowledgments 352

Teollisuusneuvos Heikki Väänänen’s fund and the North Karelia regional foundation of the 353

Finnish Cultural Foundation are acknowledged for their grants to support this work as is the 354

FORES graduate school.

355

Notes 356

The authors declare that there are no competing financial interests concerning this work.

357

References 358

[1] C. Thomsen, E. Lundanes, G. Becher. Brominated flame retardants in plasma samples from three different

359

occupational groups in Norway. J. Environ. Monit. 3 (4) (2001) 366–370.

360

[2] C. Thomsen, H. Leknes, E. Lundanes, G. Becher. A new method for determination of halogenated flame-

361

retardants in human milk using solid-phase extraction. J. Anal. Toxicol. 26 (3) (2002) 129–137.

362

[3] I. Watanabe, S. Sakai. Environmental release and behavior of brominated flame-retardants. Environ. Int. 29

363

(6) (2003) 665–682.

364

[4] M. Osako, Y. Kim, S. Sakai. Leaching of brominated flame-retardants in leachate from landfills in japan.

365

Chemosphere 57 (10) (2004) 1571–1579.

366

(21)

19

[5] T. Debenest, F. Gagné, A. Petit, C. André, M. Kohli, C. Blaise. Ecotoxicity of a brominated flame retardant

367

(tetrabromobisphenol A) and its derivatives to aquatic organisms. Comp. Biochem. Physiol. C Toxicol.

368

Pharmacol. 152 (4) (2010) 407–412.

369

[6] S.T. McKenna, T.R. Hull. The fire toxicity of polyurethane foams. Fire Sci. Rev. 5 (3) (2016) pp. 27.

370

[7] J. Jiang, J. Li, J. Hu, D. Fan. Effect of nitrogen phosphorus flame retardants on thermal degradation of wood.

371

Constr. Build. Mater. 24 (12) (2010) 2633–2637.

372

[8] Y. Xie, N. Liu, Q. Wang, Z. Xiao, F. Wang, Y. Zhang, H. Militz. Combustion behavior of oak wood (Quercus

373

mongolica L.) modified by 1,3-dimethylol-4,5-dihydroxyethyleneurea (DMDHEU). Holzforschung 68 (8) (2014)

374

881–887.

375

[9] L. Liu, M. Qian, P. Song, G. Huang, Y. Yu, S. Fu. Fabrication of green lignin-based flame retardants for

376

enhancing the thermal and fire retardancy properties of Polypropylene/Wood composites. ACS

377

Sustain. Chem. Eng. 4 (4) (2016) 2422–2431.

378

[10] E.N. Kalali, L. Zhang, M.E. Shabestari, J. Croyal, D. Wang. Flame-retardant wood polymer composites

379

(WPCs) as potential fire safe bio-based materials for building products: Preparation, flammability and

380

mechanical properties. Fire Saf. J. (2017) 10.1016/j.firesaf.2017.11.001

381

[11] F. Carosio, F. Cuttica, L. Medina, L.A. Berglund. Clay nanopaper as multifunctional brick and mortar fire

382

protection coating-wood case study. Mater. Des. 93 (2016) 357–363.

383

[12] W. Chang, Y. Pan, C. Chuang, J. Guo, S. Chen, C. Wang, K. Hsieh. Fabrication and characterization of

384

waterborne polyurethane (WPU) with aluminum trihydroxide (ATH) and mica as flame retardants. J. Polym.

385

Res. 22 (12) (2015) 1–9.

386

[13] Q.F. Sun, Y. Lu, Y.Z. Xia, D.J. Yang, J. Li, Y.X. Liu. Flame retardancy of wood treated by TiO2/ZnO coating.

387

Surf. Eng. 28 (8) (2012) 555–559.

388

[14] S.P. Kumar, S. Takamori, H. Araki, S. Kuroda. Flame retardancy of clay-sodium silicate composite coatings

389

on wood for construction purposes. RSC Adv. 5 (43) (2015) 34109–34116. [15] S. Trey, S. Jafarzadeh, M.

390

Johansson. In situ polymerization of polyaniline in wood veneers. ACS Appl. Mater. Interfaces 4 (3) (2012)

391

1760–1769.

392

[16] J. Pan, J. Mu, Z. Wu, X. Zhang. Effect of nitrogen-phosphorus fire retardant blended with Mg(OH)2/Al(OH)3

393

and nano-SiO2 on fire retardant behavior and hygroscopicity of poplar. Fire Mater. 38 (5) (2014) 817–826.

394

(22)

20

[17] V. Merk, M. Chanana, S. Gaan, I. Burgert. Mineralization of wood by calcium carbonate insertion for

395

improved flame retardancy. Holzforschung 70 (9) (2016) 867–876.

396

[18] V. Merk, M. Chanana, T. Keplinger, S. Gaan, I. Burgert. Hybrid wood materials with improved fire retardance

397

by bio-inspired mineralisation on the nano- and submicron level. Green Chem. 17 (3) (2015) 1423–1428.

398

[19] Z. Xiao, J. Xu, C. Mai, H. Militz, Q. Wang, Y. Xie. Combustion behavior of Scots pine (Pinus sylvestris L.)

399

sapwood treated with a dispersion of aluminum oxychloride-modified silica. Holzforschung 70 (12) (2016)

400

1165–1173.

401

[20] G. Chen, Y. Hu, F. Peng, J. Bian, M. Li, C. Yao, R. Sun. Fabrication of strong nanocomposite films with

402

renewable forestry waste/montmorillonite/reduction of graphene oxide for fire retardant. Chem. Eng. J. 337

403

(2018) 436–445.

404

[21] R.-N. Amiri, T. Tirri, C.-E. Wilen. Flame retardant polyurethane nanocomposite: Study of clay dispersion and

405

its synergistic effect with dolomite. J. Appl. Polym. Sci. 129 (4) (2013) 1678–1685.

406

[22] T. Tirri, M. Aubert, W. Pawelec, A. Holappa, C.-E. Wilén. Structure-property studies on a new family of

407

halogen free flame retardants based on sulfenamide and related structures. Polymers 8 (10) (2016) 360–370.

408

[23] F. Carosio, J. Kochumalayil, F. Cuttica, G. Camino, L. Berglund. Oriented clay nanopaper from biobased

409

components - mechanisms for superior fire protection properties. ACS Appl. Mater. Interfaces 7 (10) (2015)

410

5847–5856.

411

[24] C. Ma, S. Qiu, B. Yu, J. Wang, C. Wang, W. Zeng, Y. Hu. Economical and environment-friendly synthesis of

412

a novel hyperbranched poly(aminomethylphosphine oxide-amine) as co-curing agent for simultaneous

413

improvement of fire safety, glass transition temperature and toughness of epoxy resins. Chem. Eng. J. 322

414

(2017) 618–631.

415

[25] Y. Shi, T. Fu, Y. Xu, D. Li, Wang X, Wang Y, Novel phosphorus-containing halogen-free ionic liquid toward fire

416

safety epoxy resin with well-balanced comprehensive performance. Chem. Eng. J. 354 (2018) 208–219.

417

[26] X. Shi, Y. Xu, J. Long, Q. Zhao, X. Ding, L. Chen, Y. Wang. Layer-by-layer assembled flame-retardant

418

architecture toward high-performance carbon fiber composite. Chem. Eng. J. 353 (2018) 550–558.

419

[27] Y. Kim, H.S. Kim, S.M. Jo, S.Y. Kim, B.J. Yang, J. Cho, S. Lee, J.E. Cha. Thermally insulating, fire-retardant,

420

smokeless and flexible polyvinylidene fluoride nanofibers filled with silica aerogels. Chem. Eng. J. 351 (2018)

421

473–481.

422

(23)

21

[28] X. Qiu, Z. Li, X. Li, Z. Zhang. Flame retardant coatings prepared using layer by layer assembly: A review.

423

Chem. Eng. J. 334 (2018) 108–122.

424

[29] J. Alongi, R.A. Carletto, F. Bosco, F. Carosio, A. Di Blasio, F. Cuttica, V. Antonucci, M. Giordano, G. Malucelli.

425

Caseins and hydrophobins as novel green flame retardants for cotton fabrics. Polym. Degrad. Stab. 99 (1)

426

(2014) 111–117.

427

[30] F. Bosco, R.A. Carletto, J. Alongi, L. Marmo, A. Di Blasio, G. Malucelli. Thermal stability and flame resistance

428

of cotton fabrics treated with whey proteins. Carbohyd. Polym. 94 (1) (2013) 372–377.

429

[31] A. Khandual. Green flame retardants for textiles. In: S.S. Muthu, M.A. Gardetti (Eds.), Green fashion.

430

Springer, Singapore, 2016, pp. 171–227.

431

[32] J. Alongi, R.A. Carletto, A. Di Blasio, F. Carosio, F. Bosco, G. Malucelli. DNA: A novel, green, natural flame

432

retardant and suppressant for cotton. J. Mat. Chem. A 1 (15) (2013) 4779–4785.

433

[33] Carosio, F., Di Blasio, A., Alongi, J., & Malucelli, G., (2013). Green DNA-based flame retardant coatings

434

assembled through layer by layer. Polymer (United Kingdom), 54(19), 5148-5153.

435

[34] X. Zhang, J. Mu, D. Chu, Y. Zhao. Synthesis of fire retardants based on N and P and poly(sodium silicate-

436

aluminum dihydrogen phosphate) (PSADP) and testing the flame-retardant properties of PSADP impregnated

437

poplar wood. Holzforschung 70 (4) (2016) 341–350.

438

[35] Z. Wu, N. Hu, Y. Wu, S. Wu, Z. Qin. The Effect of Ultrafine Magnesium hydroxide on the tensile properties

439

and flame retardancy of wood plastic composites. J. Nanomater. 2014, ID 945308, 8 p.

440

[36] K. Nakamoto. Infrared and Raman spectra of inorganic and coordination compounds: Part B: Applications in

441

coordination, organometallic, and bioinorganic chemistry. 6th ed., John Wiley & Sons, Inc., Hoboken, New

442

Jersey, USA, 2008.

443

[37] J. Kong, S. Yu. Fourier transform infrared spectroscopic analysis of protein secondary structures. Acta

444

Biochim. Biophys. Sin. 39 (8) (2007) 549–559.

445

[38] N.B. Colthup, L.H. Daly, S.E. Wiberley. Introduction to Infrared and Raman Spectroscopy, 2nd ed. Academic

446

Press, New York, 1975.

447

[39] K. Serec, S.D. Babić, R. Podgornik, S. Tomić. Effect of magnesium ions on the structure of DNA thin films: An

448

infrared spectroscopy study. Nucleic Acids Res. 44 (17) (2016) 8456–8464.

449

[40] F. Shafizadeh. Thermal uses and properties of carbohydrates and lignins. Academic Press, 1976, 1–17.

450

[41] F. Shafizadeh, Y. Sekiguchi. Development of aromaticity in cellulosic chars. Carbon 21 (5) (1983) 511–516.

451

(24)

22

[42] M.A. Hassan, R. Kozlowski, B. Obidzinski. New fire-protective intumescent coatings for wood. J. Appl. Polym.

452

Sci. 110 (1) (2008) 83–90.

453

[43] S. Fang, Y. Hu, L. Song, J. Zhan, Q. He. Mechanical properties, fire performance and thermal stability of

454

magnesium hydroxide sulfate hydrate whiskers flame retardant silicone rubber. J. Mater. Sci. 43 (3) (2008)

455

1057–1062.

456

[44] M.A. Bahattab, J. Mosnáček, A.A. Basfar, T.M. Shukri. Cross-linked poly(ethylene vinyl acetate) (EVA)/low

457

density polyethylene (LDPE)/metal hydroxides composites for wire and cable applications. Polym. Bull. 64 (6)

458

(2010) 569–580.

459

[45] Tang S., Yang J.J. (2013) Magnesium Binding Sites in Proteins. Kretsinger R.H., Uversky V.N., Permyakov

460

E.A. (eds) Encyclopedia of Metalloproteins. Springer, New York, NY.

461

[46] Oh, H. E., & Deeth, H. C. (2017). Magnesium in milk. International Dairy Journal, 71, 89-97.

462

[47] Cuomo, F., Ceglie, A., & Lopez, F. (2011). Temperature dependence of calcium and magnesium induced

463

caseinate precipitation in H2O and D2O. Food Chemistry, 126(1), 8–14.

464

(25)

Declaration of interests

☒The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests:

Viittaukset

LIITTYVÄT TIEDOSTOT

On a heavy clay soil potassium fertilization has even increased the calcium and magnesium contents of timothy (SAARELA et ai. 1981)..

At about pH(CaCl 2 ) 6.5 the ryegrass took up the same amount (mg/pot) of magnesium from both dolomitic limestones and magnesium sulphate, despite the fact that there was a

The apparent recovery of fertilizer magnesium on mull soil and on peat soil was higher at the large potassium fertilizer level in the field experiments than at the small one

Almost all the fertilizer magnesium (4 mg/100 g soil) was extractable in ammonium acetate in unlimed soils, yet the magnesium content of heavy clay increased more than by the

As was the case with the grains, the potassium content of the straws was also in the second year significantly higher with the high nitrogen rate than the potassium con- tent of

The dependence of the magnesium content of the potato leaves, clover and timothy yields on the magnesium content of the soil and its saturation percentage, as well as on the

On the other hand, it has been proved in practice that when the cows have been given relatively big rations of magnesium, no signs of zinc deficiency have appeared if the cows have

35 peat lands which represented 6 types of these bogs were analyzed for pH, ash content, total nitrogen and phosphorus content, and for their content of calcium, magnesium,