You are on page 1of 28

Author’s Accepted Manuscript

Synthesis and properties of polyurethane wood


adhesives derived from crude glycerol-based
polyols

Shaoqing Cui, Xiaolan Luo, Yebo Li

www.elsevier.com/locate/ijadhadh

PII: S0143-7496(17)30136-7
DOI: http://dx.doi.org/10.1016/j.ijadhadh.2017.04.008
Reference: JAAD2038
To appear in: International Journal of Adhesion and Adhesives
Received date: 17 October 2016
Accepted date: 12 April 2017
Cite this article as: Shaoqing Cui, Xiaolan Luo and Yebo Li, Synthesis and
properties of polyurethane wood adhesives derived from crude glycerol-based
p o l y o l s , International Journal of Adhesion and Adhesives,
http://dx.doi.org/10.1016/j.ijadhadh.2017.04.008
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
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.
Synthesis and properties of polyurethane wood adhesives derived from

crude glycerol-based polyols

Shaoqing Cui, Xiaolan Luo, Yebo Li*

Department of Food, Agricultural, and Biological Engineering, The Ohio State University /Ohio

Agricultural Research and Development Center, 1680 Madison Ave, Wooster, OH 44691-4096, USA

* Corresponding Author. Tel.: + 1 330 263 3855; Fax: + 1 330 263 3670. E-mail address:

li.851@osu.edu (Y. Li); li.851@gmail.com.

1
Abstract

Crude glycerol, a waste stream of the biodiesel production process, is low-cost renewable

feedstock for the production of chemicals and polymers. In this study, polyurethane (PU)

adhesives were synthesized from crude glycerol-based polyols (CG-based polyols) for wood

bonding applications. Effects of different variables, including hydroxyl values of CG-based

polyols, chain extenders, and the molar ratio of NCO/OH on the properties of PU adhesives

were investigated. The chemical structures of PU adhesives were characterized, and their

thermal, mechanical, and chemical resistance properties were evaluated. The experimental

results indicated that an increase of the NCO/OH molar ratio (1.3) substantially improved

bonding strength by up to 38 MPa. Higher thermal stability and stronger chemical resistance

to hot and cold water and to alkali and acid solutions were observed comparing to vegetable

oil-based adhesives. However, the effect of the hydroxyl value of polyols on bonding strength

was not significant. Additionally, bond strength of crude glycerol-based PU adhesives was

comparable to that of some commercial PU wood adhesives. All these properties

demonstrated the potential of CG for PU wood adhesive applications, particularly for

fast-curing uses.

Keywords: Crude glycerol, Bio-polyol, Wood adhesive, Polyurethane

2
1. Introduction

Urea-formaldehyde and phenol-formaldehyde based adhesives have been widely used as

wood panel adhesives, foundry sand binders, and for the bonding of papers (Pizzi, 2015).

However, they are easily hydrolyzed and, under cyclic moisture or warm and humid

conditions, can undergo stress scissions (Robert, et al. 1994; River, et al. 1994). Moreover,

the release of formaldehyde has posed serious environmental and health concerns. To address

these problems, polyurethanes (PUs) have been widely introduced for adhesive production.

PUs are versatile polymeric materials that demonstrate excellent flexibility, strong adhesion,

high performance at low-temperature, and good stability (De Gray, 1998; John and Joseph,

1998; Malavasic, et al. 1992; Zhang, et al. 2014; Zhang, et al. 2017). In general, PUs are

produced by the reaction of polyols and isocyanates, most of which are petroleum-based.

Concerns over dwindling petroleum reserves have led to investigations into substitutes that

are produced from renewable sources.

Because of the limited choice of isocyanates, a majority of the research on renewable

substitutes used for PU production has focused on the polyol component. Natural oils

(Abraham, et al. 2007; Shah, et al. 2001; Roh, et al. 2008), lignocellulosic biomass (Nadi,

2005; Brioner, 2011), and carbohydrates such as starch (Menezes, et al. 2007) have been

widely used for the production of bio-based polyols. Specifically, vegetable oil and starch

derived polyols have been successfully used to synthesize PU adhesives. Studies reported that

the adhesion strength of PU adhesives derived from castor oil-based polyols was ten times

higher than that of commercially available adhesives, when the PU adhesives were applied to

wood joints (Keyur, et al. 2003; Desai, et al. 2003). PU adhesives produced from canola

3
oil-based polyols also demonstrated lap shear strengths superior to those of three commercial

PU adhesives (Kong, et al. 2011). Natural adhesives derived from wheat and oil palm starch

also have been demonstrated to meet industrial standards (Japanese) in terms of mechanical

properties (Kushairi, et al. 2015).

In recent years, there has been increasing interest in value-added processing of crude glycerol

into chemicals and polymers (Luo, et al. 2016; Pagliaro, et al. 2007; Tan, et al. 2013). Crude

glycerol is a low-value byproduct which is primarily obtained from the biodiesel production

process (Luo, et al. 2016). It has potential as a renewable substitute for petroleum-based

feedstocks. Studies have reported the successful production of CG-based polyols with

properties suitable for PU applications (Luo, et al. 2014; Luo, et al. 2013). Crude glycerol

impurities, such as fatty acids and/or fatty acid methyl esters, were found to be critical for the

production of high quality polyols and PU foams and coatings (Hu, et al. 2014). Compared to

vegetable oils, crude glycerol is an inexpensive byproduct and does not compete directly with

food supplies. Therefore, the use of crude glycerol in the production of bio-based PU

adhesives is expected.

In this study, the feasibility of crude glycerol-based polyols for the production of PU

adhesives for wood bonding was investigated. The effects of parameters, including hydroxyl

number of CG-based polyols, chain extenders, and the molar ratio of NCO/OH on PU

adhesive properties, were studied and the optimized parameters were determined. In addition,

the properties of PU adhesives from CG-based polyols were compared with some commercial

wood adhesives to investigate their potential applications for wood bonding.

4
2. Method and materials

2.1 Materials

Four CG-based polyols with hydroxyl numbers of 486, 422, 296 and 191 mg KOH/g, and

acid numbers below 3 mg KOH/g, were obtained from Bio100 Technologies, LLC

(Mansfield, OH). The characteristics of CG-based polyols involving the main composition,

averaged molecular weight and functionality were supplemented and are listed in Table 1 and

their representative schematics also shown in Fig.1 in the supplementary file. Polyethylene

glycerol with an average molecular weight of 400 (PEG 400), 1, 4-butanediol, ethylene

glycol, and neopentyl glycol were purchased from Fisher Scientific (Pittsburgh, PA).

Polymeric methylene-4-4’-diphenyl isocyanate (pMDI) was obtained from Covestro Bayer

Materials Science (Pittsburg, PA). Dibutyltin dilaurate (DBTDL) was obtained from Pfaltz &

Bauer (Waterbury, CT). The commercial PU adhesives used for property comparison were

Vibrs-tite PU adhesive (ND Industries, Inc., Clawson, MI), BOSTIK’S BEST wood flooring

urethane adhesive (Bostik, Inc., Middleton, MA), and waterproof Franklin Titebond PU glue

(Franklin International, Columbus, OH).

2.2 Adhesive preparation

The preparation of PU adhesives was carried out at room temperature under nitrogen

protection by mixing CG-based polyols with pMDI in the presence/absence of chain

extenders using DBTDL as a catalyst (0.1 wt% based on the total weight of polyols and

pMDI). Tables 1-3 show the formulations for the synthesis of PU adhesives from CG-based

polyols with different hydroxyl numbers (Table 1), PU adhesives with different chain

extenders from CG-based polyols with a hydroxyl number of 296 mg KOH/g (Table 2), and

5
PU adhesives from CG-based polyols with a hydroxyl number of 296 mg KOH/g with

different molar ratios of NCO/OH (Table 3).

Table 1 Formulation for the synthesis of PU adhesives from CG-based polyols with different
hydroxyl numbers
Polyols pMDI Molar ratio
Adhesives Hydroxyl number mass (g) mass (g) of NCO/OH
(mg KOH/g)
CGPU-191-1.3 191 5 2.99 1.3
CGPU-296-1.3 296 5 4.64 1.3
CGPU-422-1.3 422 5 7.38 1.3
CGPU-486-1.3 486 5 7.63 1.3

Table 2 Formulation for the synthesis of CG-based PU adhesives with different chain
extenders
Adhesives Chain extender Chain extender Polyola mass pMDI
mass (g) (g) mass (g)
CGPU-BD-1.3 1,4-butanediol 0.5 5 4.65
CGPU-EG-1.3 Ethylene glycerol 0.5 5 4.65
CGPU-NG-1.3 Neopentyl glycerol 0.5 5 4.65
CGPU-1.3 -- -- 5 4.65
Note: a Polyol hydroxyl number: 296 mg KOH/g, and molar ratio of (NCO/OH) is 1.3.
Table 3 Formulation for the synthesis of CG-based and PEG400-based PU adhesives with
different NCO/OH molar ratios
Adhesives NCO/OH Polyol mass (g) pMDI mass (g)

6
CG based / PEG400 based Based on CG polyols / PEG400
CGPU-296-1.0/ PEG400PU-1.0 1.0 6 4.30 / 4.36

CGPU-296-1.1 / PEG400PU-1.1 1.1 6 4.72 / 4.47

CGPU-296-1.2 / PEG400PU-1.2 1.2 6 5.16 / 5.15

CGPU-296-1.3 / PEG400PU-1.3 1.3 6 5.56 / 5.54

CGPU-296-1.4 / PEG400PU-1.4 1.4 6 6.08 / 5.94

CGPU-296-1.5 / PEG400PU-1.5 1.5 6 6.44 / 6.34

CGPU-296-1.6 / PEG400PU-1.6 1.6 6 6.87 / 6.73

CGPU-296-1.7 / PEG400PU-1.7 1.7 6 7.32/ 7.16

The hydroxyl number of CG-based polyols varied from 191 to 486 mg KOH/g. The NCO/OH

molar ratio varied from 1.0 to 1.7 with an interval of 0.1. The effects of chain extenders,

including neopentyl glycol, ethylene glycerol and 1,4-butanediol, on PU adhesive properties

were investigated. PU adhesives from PEG400 were also synthesized in order to investigate

property differences in comparison with CG-based PU adhesives.

2.3 Wood specimen preparation, bonding and testing

Wood pieces were cut into 100 mm×25 mm×3 mm strips and polished using sandpaper

before use. The synthesized PU adhesive was applied to the surface of two pieces of wood

strips at a thickness of 0.1 mm and a lap joint of 25 m×30 m. A load of 1 kg was placed on

top of the lap joint and left for 5 min. After that, the joined wood pieces were kept at room

temperature and at humidity of 75±5% for 7 days. The lap joint shear strength of joined wood

specimens were measured using a universal testing machine with a stretch rate of 50mm/min,

Instron 3366 (Instron Corp., Norwood, MA), in accordance with ASTM D 906. The average

value of three replicates for each sample was reported. Green strength is one of important

indexes to evaluate adhesive properties. It is closely related to the ability of an adhesive to

7
hold the substrate before reaching its ultimate bond strength when completely cured.

Regarding green strength measurement, joined wood specimens were cured at room

temperature and then directly subjected to lap shear tests at daily intervals up to 7 days. The

curing time was determined starting from CG-based polyol and isocyanate being bonded

together and ending with the surface of the adhesives being firm.

2.4 Chemical resistance

The chemical resistance of CG-based PU adhesives were tested under four conditions: cold

water, boiling water, acid solution, and alkali solution. Wood specimens bonded with PU

adhesives were immersed in cold water at 4 oC and hot water at 100 oC for 1 day, and in acid

( hydrochloric acid, pH 2) and alkaline (sodium hydroxide, pH 10) solutions at 80 oC for 1

hour. After that, the specimens were dried at room temperature for 7 days and then subjected

to lab shear strength tests as described above.

2.5 FT-IR analysis

Fourier transform-infrared (FT-IR) spectra of CG-based polyols and the resulting PU

adhesives were obtained on a Spectrum Two IR spectrometer (PerkinElmer, Waltham, MA)

with 32 scans at a resolution of 2 cm-1.

2.6 Thermogravimetric analysis

Thermogravimetric analysis (TGA) was performed using a Q50 thermogravimeter (TA

Instruments, New Castle, DE) by heating PU adhesive samples from 50 to 600 oC at a rate of

10 oC/ min under a nitrogen atmosphere.

8
3. Results and discussion

3.1 Effect of the hydroxyl number of polyols

Table 4 shows the effects of the hydroxyl number of CG-based polyols on lap shear strength

and curing time of PU adhesives. As the hydroxyl number increased from 191 to 486 mg

KOH/g, the average lap shear strength of the CG-based PU adhesives gradually increased

from 31.6 to 37.1 MPa, while curing time decreased from approximately 5 to 2.5 min. These

results were mainly due to the increased crosslinking density of PU adhesives. Under the

same NCO/OH molar ratio, as the hydroxyl number of polyols increased, the content of

hydroxyl groups increased, causing more isocyanates to participate in the reaction. This

resulted in the formation of a more compact crosslinking structure in the PU adhesives.

Similar results also have been observed for castor oil-based PU adhesives (Keyur, et al. 2003;

Moghadam, et al. 2003). Compared to castor oil-based PU adhesives and other vegetable

oil-based PU adhesives (Ang, et al. 2014; Li and Li, 2014), CG-based PU adhesives showed

shorter curing times. This might be caused by the low molecular weight of the CG polyols

and the residual glycerol, which plays a role as chain extender (Luo, et al. 2013; Hu and Li,

2015). As demonstrated in Table 1 of the supplementary file, the free glycerol content of the

employed CG-based glycerol ranges from 5.7% to 20.9% and molecular weight from 932 to

669. Besides, with an increasing hydroxyl number, the hard segment of the obtained

adhesives increased from 37.9% to 60.8%, which further confirmed the results of increasing

lap shear strength and decreasing curing time. The results in Table 4 also indicate that the

hydroxyl number of the CG-based polyols had more obvious effects on curing time than on

adhesion strength. Considering a balance between lap shear strength and curing time, the

9
CG-based polyol with a hydroxyl number of 296 mg KOH/g was chosen as a suitable polyol

for further studies.

Table 4 Effects of the hydroxyl number of CG-based polyols on lap shear strength and curing
time of PU adhesives
Adhesives Hydroxyl number of CG Lap shear strength Curing time Hard segment
polyol (mg KOH/g) (MPa) (min) (wt %)
CGPU-191-1.3 191 31.6±3.3 4.0~5.0 37.9
CGPU-296-1.3 296 35.8±4.3 3.5~4.5 48.6
CGPU-422-1.3 422 36.3±2.7 3.5~4.0 57.4
CGPU-486-1.3 486 37.1±2.0 2.5~3.5 60.8

3.2 Effect of chain extenders

The effect of chain extenders on lap shear strength of CG-based PU adhesives are shown in

Table 5. No obvious change was observed with the addition of a chain extender. PU

adhesives containing 1, 4-butanediol and neopentyl glycol resulted in a lower lap shear

strength than that of an analogue with ethylene glycol as a chain extender. This was most

likely because both 1, 4-butandiol and neopentyl glycol had a longer chain length than that of

ethylene glycol, thus improving the flexibility of PU structures. In addition, the presence of

two side methyl groups of neopentyl glycol had a plasticizing effect on PU performance to

some extent, reducing the lap shear strength of its PU adhesives. Compared to PU adhesives

with the addition of chain extenders, CG-based PU adhesives without chain extenders

exhibited the highest lap shear strength. This was mainly due to the presence of residual

glycerol (around 6.5%) in CG-based polyols (Luo, et al. 2013; Hu and Li, 2014; Hu, et al.

2015), resulting in the formation of a crosslinked structure and further increasing the lap

shear strength of the PU adhesives.

10
Table 5 Lap shear strength of CG-based PU adhesives with the addition of different chain
extenders
Samplea Chain Extender Lap shear strength (MPa)
CG-BD-1.3 1,4-butanediol 35.7±3.3
CG-EG-1.3 Ethylene glycerol 33.5±2.7
CG-NG-1.3 Neopentyl glycerol 34.5±1.8

CG-1.3 None 35.8±4.3


a
Note: The experiments were carried out with NCO/OH of 1.3 and hard segment around 49.1%.

3.3 Effect of NCO/OH ratio

In this study, the effects of the NCO/OH molar ratio on the properties of CG-based PU

adhesives that were produced by the reaction between CG-based polyols with a hydroxyl

number of 296 mg KOH/g and pMDI, were investigated. In order to examine the property

differences in PU adhesives from different structures of polyols, PU adhesives from PEG400,

which has an approximate hydroxyl number of 280 mg KOH/g, were also produced with the

same NCO/OH molar ratios, from 1.0 to 1.7. Table 6 presents the properties of PU adhesives

from both polyols under the varying NCO/OH molar ratios. An obvious increase in bond

strength with both CG-based and PEG400-based PU adhesives was observed as the NCO/OH

molar ratio incrementally increased from 1.1 to 1.3. However, with a further increase of the

NCO/OH molar ratio from 1.4 to 1.7, the bond strength of both PU adhesives gradually

decreased. An optimal NCO/OH molar ratio for the highest shear strength of both PU

adhesives was 1.3. As the NCO/OH molar ratio increased, PU adhesives with higher crosslink

densities were achieved as demonstrated with increasing hard segment contents, resulting in

increased rigidity of the adhesive. Moreover, the increased NCO/OH molar ratio provided

more free isocyanates, which can react with active groups on the surface of wood samples,

and further strengthen the adhesive bond (Keyur, et al. 2003). However, when beyond a

11
critical ratio (1.3), an increase in the NCO/OH molar ratio will likely cause more complex

side reactions, such as the reaction of isocyanate groups with urethanes to form allophanates,

the reaction of isocyanate groups with water present in wood to form ureas, and the reaction

of ureas with isocyanate groups to form biuret, which increased the stiffness of PU adhesives

and resulted in decreased adhesion strength (Moghadam, et al. 2016; Kong, et al, 2011).

Additionally, increased hard segment content might enhance the rigidity of a bonded joint

which, in contrast, may reduce the bond strength. Similar changes in adhesion strength with

an increase in the NCO/OH molar ratio have been reported for PU adhesives from ricinoleic

acid-based polyols (Moghadam, et al. 2016; Kong, et al. 2011; Saetung, et al. 2015). As the

NCO/OH molar ratio increased from 1.0 to 1.7, the curing time of PU adhesives from

CG-based polyols and PEG400 was reduced from 11 to 3 min and from 65 to 45 min,

respectively. Compared to PEG-based PU adhesives, CG derived adhesives showed much

higher lap shear strengths but shorter curing times. This was mainly ascribed to structural

differences between the two polyols and the residual glycerol in CG-based polyols. CG-based

polyols are mainly composed of monoglycerides, diglycerides, and glycerol, whose hydroxyl

groups are similar to PEG400. In contrast, PEG400 is a linear polyether with two terminal

hydroxyl groups. It is more flexible than CG-based polyols, therefore, lower adhesion

strength was obtained. As mentioned in section 3.2, the residual glycerol could act as a

cross-linker to react with isocyanates for the formation of a compact network structure,

resulting in high adhesion strength of the PU adhesive. It was concluded that both

cross-linking density and network structure affected the adhesive properties to some degree.

12
Table 6 Lap shear strength and curing time of designed PU adhesives based on different
NCO/OH molar ratios
CG-based PU Lap shear Curing Hard PEG400-based Lap shear Curing Hard
a b
adhesives strength time Segment PU adhesives strength time (min) segment
(MPa) (min) (wt %) (MPa) (wt %)
CGPU-296-1.0 16.6±2.3 10~11 42.1 PEG400PU-1.0 0.8±1.3 60~65 40.7
CGPU-296-1.1 31.6±3.3 7~7.5 44.4 PEG400PU-1.1 2.3±1.7 60~65 43.1
CGPU-296-1.2 33.2±2.7 5~5.2 46.6 PEG400PU-1.2 12.2±2.1 55~60 45.2
CGPU-296-1.3 36.8±3.5 4~4.5 48.6 PEG400PU-1.3 17.4±2.5 55~60 47.2
CGPU-296-1.4 35.8±3.1 3~3.6 50.4 PEG400PU-1.4 17.1±2.3 50~55 49.0
CGPU-296-1.5 33.9±2.4 3~3.7 52.1 PEG400PU-1.5 16.4±1.5 50~55 50.8
CGPU-296-1.6 31.6±3.5 3~3.5 53.8 PEG400PU-1.6 16.1±1.6 45~50 52.4
CGPU-296-1.7 31.0±3.5 3~3.5 55.3 PEG400PU-1.7 15.5±1.4 45~50 53.9
Notes: a 296 is the hydroxyl number of the CG polyol. Numbers of 1.0-1.7 are the NCO/OH molar ratios
used in the formulations.
b
1.0-1.7 are the NCO/OH molar ratios used in the formulations.

3.4 Properties of adhesives derived from blended polyols

Blends of different ratios of CG-based polyols and PEG400 were investigated to produce PU

adhesives with improved curing times using an NCO/OH molar ratio of 1.3. Table 7 shows

the formulations for the production of PU adhesives from the blended polyols (CG-based

polyols and PEG400) and their adhesion properties. The strength of hybrid adhesives were

greatly improved compared to PEG400-based adhesives, and their curing times were

significantly reduced from 60 min to 6~8 min. This significant change in adhesive

performance was probably ascribed to free glycerol present in the CG-based polyols. In order

to examine the effect of glycerol on the properties of PEG400-based PU adhesives, mixtures

of pure glycerol and PEG400 were employed for the production of PU adhesives. The

adhesive properties are shown in Table 8. As the content of pure glycerol in PEG400

increased from 0 to 40%, the curing time significantly decreased from 60 to 15 min. With a

further addition of glycerol to PEG 400 (80% glycerol), the curing time decreased to 6 min.

13
Therefore, the removal of glycerol in CG-based polyols was required to improve its curing

time.

Table 7 Formulations for the production of PU adhesives from blended polyols and their
adhesion properties
Adhesives CG polyol content Mass of CG Mass of Mass of Lap shear Curing
samples*a in the blend (w/w) polyol (g) PEG400 (g) pMDI (g) strength (MPa) time (min)
PU-0-100 0% 0 4.0 3.41 17.4±2.5 55~60
PU-10-90 10% 0.4 3.6 3.47 25.4±1.7 20~25
PU-20-80 20% 0.8 3.2 3.50 29.0±1.9 18~22
PU-30-70 30% 1.2 2.8 3.53 30.9±1.2 15~17
PU-40-60 40% 1.6 2.4 3.56 31.2±1.5 13~15
PU-50-50 50% 2.0 2.0 3.58 33.2±2.4 8~10
PU-60-40 60% 2.4 1.6 3.61 33.5±2.1 8~10
PU-70-30 70% 2.8 1.2 3.64 33.6±2.1 6~8
PU-80-20 80% 3.2 0.8 3.67 34.3±2.6 6~8
PU-90-10 90% 3.6 0.4 3.70 35.0±1.9 4~5.5
PU-100-0 100% 4.0 0 3.73 35.8±2.7 4~5.5
a
Note: The molar ratio of NCO/OH is 1.3.

Table 8 Adhesive properties based on PEG-400 with different contents of glycerol

Pure glycerol content Lap shear strength Curing time


(w/w) (MPa) (min)
0 17.3 55~60
40% 18.1 15~20
80% 18.4 6~10
* Formulation of this designed experiment: PEG-400 4g, pMDI 3.4g, molar ratio of –NCO/OH 1.3, and
different contents of pure glycerol.

3.5 Green strength

In this study, the curing performance of PU adhesives from CG-based polyols with NCO/OH

molar ratios from 1.2 to 1.4, were investigated by measuring lap shear strength over one week.

It can be seen from Fig.1 that the adhesion strength of three CG-based PU adhesives

increased markedly during the first 4 days, and then increased slightly until day 6. Almost no

significant change in adhesion strength was observed between days 6 and 7 of testing.

Approximately 57% and 86% of the maximum strength were obtained after curing for 2 and

14
4 days, respectively. This suggested that the curing reaction occurred quickly during the first

4 days and then slowed down until full curing was achieved after 6 days. As for individual

adhesive samples, the adhesion strength increased as the NCO/OH ratio increased from 1.2 to

1.3 and then decreased with a further increase of the NCO/OH ratio to 1.4. This could be

explained by the occurrence of more complex side reactions which was discussed in section

3.3. Compared with some reported vegetable oil-based PU adhesives, CG-based PU

adhesives showed a shorter curing time when their bond strength reached 85% of the

maximum strength. They also exhibited higher bond strength after 4-day curing (Keyur, et al.

2003; Kong, et al. 2011).

3.6 Chemical resistance

The chemical resistance of CG-based PU adhesives, with a NCO/OH molar ratio of 1.3, to

cold water, hot water, and acid and alkali solutions is shown in Fig. 2. GC-based PU

adhesives showed superior resistance to cold water. Hot water, acid solution, and alkali

solution had negative effects on the adhesion strength. It was also observed that acid and

alkali conditions caused stronger hydrolysis of PU adhesives than hot water. Adhesives fared

the worst in alkali conditions, and showed the lowest lap shear strength compared to other

conditions. These results are in agreement with previous studies on the chemical resistance of

biomass-based PU adhesives (Desai, et al. 2003). The sharp reduction in lap shear strength of

CG-based PU adhesives after alkali treatment was probably related to the degradation of the

wood surface due to penetration of the alkali solution, thereby deteriorating the contact

between the wood and adhesive and loosening the cured bond (Desai, et al. 2003; Zanuttini,

et al. 1999; Tamburini, 1970;).

15
3.7 Infrared spectroscopy

Fig. 3 shows the FTIR spectra of a CG-based polyol with a hydroxyl number of 296 mg

KOH/g and its corresponding PU adhesive with an NCO/OH molar ratio of 1.3. For the

-1
CG-based polyol, a strong broad stretching band at 3445 cm due to hydroxyl groups, a

stretching band at 1745 cm-1 due to carbonyl groups of esters, and a stretching peak at 1100

cm -1 due to C-O bonds were observed in Fig. 3(a). After the reaction of the CG-based polyol

with isocyanates, the produced CG-based PU adhesive showed some obvious changes in its

FT-IR spectrum (Fig. 3(b)). A strong absorption band at 3340 cm -1 characteristic of the N-H

group and an absorption band at around 1736 cm -1 due to carbonyl groups of urethanes were

observed, demonstrating the presence of urethane structures in the adhesive. Besides, a clear

strong absorption band at 2274 cm-1 was also observed in Fig. 3(b), which indicates the

presence of excess –NCO groups in the PU adhesive. The appearance of expected

characteristic structures in this PU adhesive confirmed the successful synthesis of PU

adhesives from the CG-based polyols.

3.8 Thermal properties

Fig. 4 shows thermogravimetric (TG) and derivative thermogravimetric (DTG) curves of a

PU adhesive produced from a CG-based polyol with a hydroxyl number of 296 mg KOH/g

and an NCO/OH molar ratio of 1.3 under a nitrogen atmosphere. Three stages of weight loss

of the CG-based PU adhesive were observed in the DTG curve of CG-based PU adhesive,

and the degradation started at about 200 oC and ended at 550 oC, similar to the degradation

behavior of vegetable oil-based PUs (Javni, et al. 2000; Ni, et al. 2010.) The first stage of

degradation occurred at temperatures from approximately 200 to 300 oC, which resulted from

16
the dissociation of urethane linkages to isocyanates and alcohols and/or to amines and olefins

with a loss of CO2 (Javni, et al. 2000). The second and third stages of degradation occurred

over the temperature range of 300 to 550 oC, and may have been due to the intensified

decomposition of urethane bonds (Lee et. al. 2007) and the scission of fatty acid chains in the

polyol structures (Luo et al. 2013; Kong, et al. 2011). Overall, the CG-based PU adhesive

showed good thermal stability compared to analogs derived from vegetable oils (Dweck, et al.

2004; Javni, et al. 2000).

3.9 Comparison with commercial wood adhesives

A PU adhesive prepared from a CG-based polyol with an NCO/OH molar ratio of 1.3 was

compared with three commercial PU wood adhesives. As shown in Table 9, CG-based PU

adhesives had lap shear strengths superior to that of the commercial PU wood adhesives,

which suggests that CG-based PU adhesives have potential for wood bonding applications.

Further investigation is needed to develop methods to prolong the curing time for use in other

applications.

Table 9 Comparison of CG-based PU adhesives and three commercial PU wood adhesives

Sample Lap shear strength (MPa)


Bostik’s Best adhesive 35.5±2.7
J.E. Moser’s Wood Glue 30.2±3.6
Sikaflex PU adhesive 37.7±2.1
CG-based PU adhesive 36.8±2.5

17
4. Conclusions

In this study, PU adhesives were successfully synthesized from CG-based polyol and pMDI.

With an increase of the hydroxyl number of CG-based polyols, the resulting PU adhesives

showed increased bond strength but faster curing time. The molar ration of NCO/OH of 1.3

was preferred in terms of adhesion strength. CG-based PU adhesives exhibited superior

resistance to cold water, moderate resistance to hot water, and relatively weak resistance to

acid and alkali solutions. Compared to some commercial PU wood adhesives, CG-based PU

adhesives demonstrated higher bond strength, suggesting they have potential for wood

bonding applications. Due to the fast curing time, the applications of CG-based PU adhesives

may be targeted at fast-curing adhesive applications.

18
Acknowledgements

This project is supported by funding from USDA-NIFA Critical Agricultural Materials

Program (No. 2012-38202-19288). The authors would like to thank Mrs. Mary Wicks

(Department of Food, Agricultural and Biological Engineering, OSU) for reading through the

manuscript and providing useful suggestions. The authors would also like to thank Josh

Borgemenke (Department of Food, Agricultural and Biological Engineering, OSU) for giving

critical suggestions.

19
References
Abraham, T.W., Carter, J.A., Malsam, J., Zlatanic, A.B., 2007. Pittsburg State University,
Oligomeric Polyols from Palm-based Oils and Polyurethane Compositions Made Therefrom,
WO2007123637.
Ang, K.P., Lee, C.S., Cheng, S.F. Chuah, C.H., 2014. Polyurethane wood adhesive from palm
oil-based polyester polyol. J. Adhes. Sci. Technol. 28, 1020-1033.
Briones, R., Serrano, L., Younes, R.B., Mondragon, I., Labidi, J., 2011. Polyol production by
chemical modification of date seeds. Ind. Crops. Prod. 34, 1035-1040.
De Gray, D., 1998. PUR adhesives offer solutions for assembly challenges. Adhes. Age. 41,
23-24.
De Menezes, A.J., Pasquini, D., Curvelo, A.A., Gandini, S., 2007. Novel thermoplastic
materials based on the outer-shell oxypropylation of corn starch ganules. Biomacromolecules.
8, 2047-2050.
Desai, S.D., Anurag, L.E., Kumar, S.V., 2003. Biomaterial Based Polyurethane Adhesive for
Bonding Rubber and Wood Joints. J. Polym. Res. 10, 275-281.
Desai, S.D., Patel, J.V., Sinha, V.K., 2003. Polyurethane adhesive system from
biomaterial-based polyol for bonding wood. Int. J. Adhes. Adhes. 23. 393-399.
Dweck, J., Sampaio, C.M.S., 2004. Analysis of the thermal decomposition of commercial
vegetable oils in air by simultaneous TG/DTA. J. Therm. Anal. Calorim. 75: 385-391.
Ebewele, R.O., River, B.H., Myers, G.E., 1994. Behavior of amine-modified
urea-formaldehyde-bonded wood joints at low formaldehyde/urea molar ratios. J. Appl.
Polym. Sci. 52, 689-700.
Hu, S., Luo, X., Li, Y. 2015. Production of polyols and waterborne polyurethane dispersions
from biodiesel-derived crude glycerol. J. Appl. Polym. Sci. 132, 41425-41433.
Hu, S., Li, Y., 2014. Polyols and polyurethane foams from acid-catalyzed biomass
liquefaction by crude glycerol: effects of crude glycerol impurities. J. Appl. Polym. Sci. 131,
9054-9062.
Javni, I., Petrovic, Z.D., Guo, A., 2000. Rachel Fuller. Thermal stability of polyurethane
based vegetables oils. J. Appl. Polym. Sci. 77, 1723-1734.
John, N., Joseph, R., 1998. Rubber solution adhesives for wood to wood bonding. J. Appl.
Polym. Sci. 68, 1185-1189.
Keyur, P.S., Sujata, S.K., Natvar, K.P., Animesh, K.R., 2003. Castor Oil based polyurethane
adhesives for wood-to-wood bonding. Int. J. Adhes. Adhes. 23, 269-275.
Kong, X.H., Liu, G.G., Curtis. J.M., 2011. Characterization of canola oil based polyurethane
wood adhesives. Int. J. Adhes. Adhes. 31,559-564.
Lee, C.S., Ooi, T.L., Chuan, C.H., 2007. Synthesis of Palm Oil-based Diethanolamides. J. Am.
Oil. Chem Soc. 84, 945-952.

20
Li, A., Li, K., 2014. Pressure-sensitive adhesives based on epoxidized soybean oil and
Dicarboxylic acids. ACS Sustainable Chem. Eng. 2, 2090-2096.
Luo, X., Ge, X., Cui, S., Li, Y., 2016, Value-added processing of crude glycerol into
chemicals and polymers. Bioresource Technol. 215, 144-154.
Luo, X., Hu, S., Zhang, X., Li, Y., 2013. Thermochemical conversion of crude glycerol to
biopolyols for the production of polyurethane foams. Bioresource Technol. 139, 323-329.
Luo, X., Li, Y., 2014. Synthesis and characterization of polyols and polyurethane foams from
PET waste and crude glycerol. J. Polym. Environ. 22, 318-328.
Malavasic, T., Cernilec, N., Mirceva, A., Osrekar, U., 1992. Synthesis and adhesive properties
of some polyurethane dispersion. Int. J. Adhes. Adhes. 112,38-42.
Moghadam, P.N., Yarmohamadi, M., Hasanzadeh, R., Nuri. S., 2016. Preparation of
Polyurethane Wood adhesives by polyols formulated with polyester polyols based on castor
oil. Int. J. Adhes. Adhes. 68, 273-282.
Munchow, E.A., Ferreira, A.C., Machado, R.M., Ramos, T.S., Rodrigues-Junior, S.A., Zanchi,
C.H., 2014. Effect of Acidic solutions on the surface degradation of a micro-hybrid
composites Resin. Braz. Den. J. 25, 321-326.
Nadji, H., Bruzzese, C., Belgacem, M.N., Benaboura, A., Gandini, A., 2005. Oxypropylation
of lignins and preparation of rigid polyurethane foams from the ensuing polyols. Macromol.
Mater. Eng. 290, 1009-1016.
Ni, B., Yang, L., Wang, C., 2010. Synthesis and thermal properties of soybean oil-based
waterborne polyurethane coatings. J. Therm. Anal. Calorim. 100: 239-246.
Pagliaro, M., Ciriminna, R., Kimura, H., Rossi, M., Pina, C.D., 2007. From Glycerol to
Value- Added Products. Angew. Chem. Int. Ed. 46, 4434-4440.
Pizzi, A., 2015. Synthetic Adhesives for wood panels: Chemistry and Technology, in: Mittal
K.L. (Eds.), Progress in adhesion and adhesives. Scrivener Publishing, Beverly, pp. 66-74.
River, B.H., Ebewele, R.O., Myers, G.E., 1994. Failure mechanisms in wood joints bonded
with urea-formaldehyde adhesives. Eur. J. Wood Prod. 52, 179-184.
Roh, Y., Kumar, R., Zhao, C. L., Kaczan, D., Smiecinski, T.M., 2008. BASF, Polyol formed
from an epoxidized Oil, WO2009138411.
Saetung, A., Tsupphayakorn-ake, P., Tulyapituk, T., Seatung, N., Phinyocheep, P., Pilard, J. F.,
2015. The chain extender content and NCO/OH ratio flexibly tune the properties of natural
rubber-based waterborne polyurethanes. J. Appl. Polym. Sci. 132, 42505 -42517.
Salleh, K.M., Hashim, R., Sulaiman, O., Hiziroglu. S., Nadhari. W.N.A.W., Karim. N.A.,
Jumhuri. N., Ang, L. Z. P., 2015. Evaluation of properties of starch-based adhesives and
particleboard manufactured from them. J. Adhes. Sci. Technol., 29, 319-336.
Shah, A.M., Shah, T.M., 2001. Process for Production Polyols and Polyols for Polyurethane,
US6258869.
Tamburini, U., 1970. Alkaline degradation of wood: Effects on Young’s Modulus. Wood Sci.

21
Technol. 4, 284-291.
Tan, H.W., Abdul Aziz, A.R., Aroua, M.K., 2013. Glycerol production and its applications as
a raw material: A review. Renew. Sust. Energ. Rev. 27, 118-127.
Zhang, L. H., Brostowitz, N.R., Cavicchi, K.A., Weiss, R.A., 2014. Perspective: Ionomer
research and applications. Macromol. React. Eng. 8, 81-99.
Zhang, L.H., Chen, L.F., Rowan, S.J., 2017. Trapping dynamic disulfide bonds in the hard
segment of thermoplastic polyurethane elastomers. Macromol. Chem. Phys. 218, 1600320
(1-10).

22
Figures and Tables

Fig. 1 Lap shear strength of CG-based PU adhesives prepared at different NCO/OH molar

ratios as a function of time

Fig. 2 Lap shear strength of CG-based PU adhesives after chemical treatment (cold water, hot

water, acid solution, and alkali solution)

Fig. 3 FTIR spectra of a CG-based polyol and its derived PU adhesive

Fig. 4 Thermogravimetric (TG) and derivative thermogravimetric (DTG) curves of a

CG-based PU adhesive

23
Figures

Fig.1 Lap shear strength of CG-based PU adhesives prepared at different NCO/OH molar
ratios as a function of time

24
Fig. 2 Lap shear strength of CG-based PU adhesive after chemical treatment (cold water, hot
water, acid solution, and alkali solution)

25
Fig.3 FTIR spectra of a CG-based polyol and its derived PU adhesive

26
Fig. 4 Thermogravimetric (TG) and derivative thermogravimetric (DTG) curves of a
CG-based PU adhesive

27

You might also like