You are on page 1of 7

Fuel Processing Technology 91 (2010) 559–565

Contents lists available at ScienceDirect

Fuel Processing Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / f u p r o c

Densification characteristics of corn cobs


Nalladurai Kaliyan, R. Vance Morey ⁎
Department of Bioproducts and Biosystems Engineering, University of Minnesota, 1390 Eckles Avenue, St. Paul, MN 55108, USA

a r t i c l e i n f o a b s t r a c t

Article history: Corn cobs are potential feedstocks for producing heat, power, fuels, and chemicals. Densification of corn cobs
Received 31 December 2008 into briquettes/pellets would improve their bulk handling, transportation, and storage properties. In this
Received in revised form 29 December 2009 study, densification characteristics of corn cobs were studied using a uniaxial piston-cylinder densification
Accepted 1 January 2010
apparatus. With a maximum compression pressure of 150 MPa, effects of particle size (0.85 and 2.81 mm),
moisture content (10 and 20% w.b.), and preheating temperature (25 and 85 °C) on the density and
Keywords:
Biomass
durability of the corn cob briquettes (with diameter of about 19.0 mm) were studied. It was found that the
Briquettes durability (measured using ASABE tumbling can method) of corn cob briquettes made at 25 °C was 0%. At
Briquetting both particle sizes, preheating of corn cob grinds with about 10% (w.b.) moisture content to 85 °C produced
Corn cobs briquettes with a unit density of > 1100 kg m-3 and durability of about 90%.
Densification © 2010 Elsevier B.V. All rights reserved.
Pellets

1. Introduction The higher heating value (HHV) of corn cobs ranges from 18.3 to
18.8 MJ/kg of dry matter [3,19]. Ebeling and Jenkins [19] reported that
Corn cobs are one of the potential agricultural biomass feedstocks the proximate analysis of corn cobs resulted in 80.10% volatiles, 1.36%
for renewable energy industries in the U.S. to abate the current energy ash, and 18.54% fixed carbon on a dry mass basis. They also reported
and the greenhouse gas problems [1]. Corn cobs can be used for that the ultimate analysis of corn cobs resulted in 46.58% carbon,
producing heat, power, gas/liquid fuels, and a wide variety of chemical 5.87% hydrogen, 45.46% oxygen, 0.47% nitrogen, 0.01% sulfur, 0.21%
products such as furfural, xylitol and activated carbon [1–8]. chlorine, and 1.40% ash on a dry mass basis [19]. Thus, corn cobs are
Crofcheck and Montross [9] found a greater yield of glucose (i.e., suitable for heating applications especially due to their low ash
ethanol) from corn cobs than other corn residues such as stalks or contents (<2% d.b.) compared to other agricultural residues [2,3,19].
leaves plus husks. The need for densification of corn cobs into briquettes/pellets
Cobs represent about 8 to 9% of the aboveground dry matter (grain could be justified based on the end use. Corn cobs are a part of the corn
plus residues) at grain physiological maturity [10,11]. The yield of stover. Corn stover includes a mixture of individual pieces of cobs,
corn cobs may range from 1.42 to 1.53 dry t/ha [12]. Currently, after husks, stalks and leaves having different shapes and sizes. Baling of
combining the grain, corn residues are collected as baled corn stover, corn stover can gather these individual pieces into a large cylindrical
which includes cobs, husks, leaves and stalks [13]. About 15 to 20% (d. compact with density of up to 150 kg m-3 (9 lb ft-3) [20]. Chopping or
b.) of aboveground corn residues (non-grain) is corn cobs [11,13,14]. grinding of corn stover could result in a relatively uniform product;
Corn cob moisture content may range from 20 to 55% (w.b.) however, chopping/grinding may not increase the bulk density
depending on the grain moisture content at the time of harvest significantly higher than the baled density. For example, fine grinding
[3,15]. With the existing corn stover collection process, most of the of corn stover using a hammer mill to a particle size of 0.34 mm
corn cobs are left on the field. Modifications to the existing corn resulted in a bulk density of 161 kg m-3 (10 lb ft-3) [21]. On the other
combining method have been suggested to collect corn cobs in a hand, corn cobs exist as individual pieces with similar shapes and
single-pass when the grain is being combined [1,12,16]. In the U.S. sizes. The bulk density of whole corn cobs would range from 160 to
Mid-West, corn cobs are collected during a limited harvest time, 210 kg m-3 (10 to 13 lb ft-3), which is higher than the density of corn
October–November. Therefore, there is a need to store corn cobs for stover bales (i.e., 150 kg m-3) [17,20]. Either in baled form or
year-around supply to large corn cob processing plants. Large volumes chopped/ground form, corn stover may be difficult to handle,
of corn cobs can be stored as piles outdoors or indoors [17,18]. transport, store and use, whereas the whole corn cobs could be
handled, transported, stored and used relatively easily compared to
the baled or chopped/ground corn stover. It appears that briquetting/
⁎ Corresponding author. Tel.: + 1 6126258775; fax: + 1 6126243005. pelleting of corn stover or corn cobs can produce uniform products
E-mail address: rvmorey@umn.edu (R.V. Morey). with bulk density of 500–600 kg m-3 (31-38 lb ft-3) [21]. Therefore,

0378-3820/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2010.01.001
560 N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565

changing the physical form of corn stover into briquettes/pellets is Table 2


essential to improve its transportation, handling, storage, and use. But, Compositions of corn cobs, corn stover, and switchgrass.

corn cobs can be used without briquetting/pelleting for majority of Component Corn cobs used in this Corn stover [21] Switchgrass [21]
applications such as industrial scale heating. For applications study (% of dry matter) (% of dry matter) (% of dry matter)
requiring high quality feedstock such as home heating, corn cobs (n = 1) (n = 1) (n = 1)

may need to be pelleted. Furthermore, although corn cobs are almost Cellulosea 40.0 49.4 43.8
cylindrical in shape, a mixture of whole and broken pieces of cobs may Hemicelluloseb 41.4 26.2 28.8
c
Lignin 5.8 8.8 9.2
result due to breakage of cobs during combining. Therefore, corn cobs
Crude protein 2.5 3.6 3.9
may lack free-flowing properties. We propose that densification of Starch 2.1 0.4 1.0
corn cobs into briquettes/pellets would result in consistent, high Crude fat 0.7 0.7 0.9
density products with uniform shapes and sizes, which can be Water soluble 1.1 7.9 2.2
efficiently handled, transported, stored, and used. carbohydrates
Moisture content 1.7 5.4 5.7
A literature review by Kaliyan and Morey [22] found that biomass
Ash 1.8 11.2 5.0
feedstock variables such as particle size, moisture content and steam a
Cellulose = acid detergent fiber – lignin.
conditioning/preheating temperature affected the density and durabil- b
Hemicellulose = neutral detergent fiber – acid detergent fiber. The hemicellulose
ity of densified products such as pellets and briquettes. They concluded content is higher than cellulose content because of the approximate estimation of
that particle sizes of geometric mean diameters of 0.5 to 1.0 mm, hemicellulose content by the difference between neutral detergent fiber and acid
moisture contents of 8 to 20% wet basis, and preheating temperatures of detergent fiber contents.
c
65 to 100 °C produced high quality (i.e., high density and durability) Lignin values measured for the biomass materials were acid insoluble lignin
contents (not total lignin contents, which would be much higher than the values
densified products for a variety of biomass feedstocks such as corn reported in Table 2 [21]).
stover, alfalfa, wheat straw, sawdust and animal feed materials.
The main objective of this research was to study the densification
characteristics of corn cobs. The specific objective was to investigate Moisture content of the corn cob grinds was determined using the
the effects of particle size, moisture content, and preheating procedure given in ASABE Standard S358.2 [23]. The moisture content
temperature on corn cob briquette density and durability at a values reported in this article are on a wet mass basis unless
compression pressure of 150 MPa. The briquetting study was mentioned otherwise. To increase the moisture content of the grinds
conducted at a compression pressure of 150 MPa because commercial to 10% or 20% (w.b.), a predetermined amount of distilled water was
roll press briquetting or pelleting machines operate at or around added to the grinds, thoroughly mixed and stored in zip-lock plastic
150 MPa pressure [22]. bags at 5 °C for 48 h for tempering. Bulk density of the corn cob grinds
was calculated from the mass of grind that occupied a 250-mL glass
container (Table 1). While measuring the bulk density of the grinds,
2. Materials and methods the glass container with the sample was tapped gently for about 4 to 6
times on a lab-bench to remove large voids inside the sample as well
2.1. Corn cob samples as to reduce the sample filling errors. Particle size and its distribution
of the corn cob grinds were determined based on ASABE Standard
The corn cob (from an unspecified variety of field corn) samples S319.3 [24] (Table 1). A corn cob grind sample was sent to a forage
used in this study were supplied by Agricultural Utilization Research analysis laboratory (Dairy One, Ithaca, NY; www.DairyOne.com) to
Institute (AURI), Waseca, MN as coarsely ground cobs. The whole corn determine the chemical composition of corn cobs based on the
cobs were collected during the fall (October) 2007 harvest season and analytical procedures given by AOAC International [25] (Table 2).
were stored at about 12–15% (w.b.) moisture content in cardboard
containers until used for this study in May 2008. At AURI, the whole 2.2. Briquetting procedure
corn cobs were ground using an 18.7 kW (25.0 hp) hammer mill
(Jacobson Quality Machinery, Minneapolis, MN) fitted with 6.4 mm Corn cob briquettes were made using a uniaxial, piston-cylinder
(1/4 in) screen. The coarsely ground corn cobs had a moisture content densification apparatus in the laboratory (Fig. 1). The details of the
of 7.5% (w.b.) due to the moisture loss from the corn cobs during the
hammer milling process (Table 1). This coarser corn cob grind served
as one of the two particle sizes tested in the study. The second particle
size tested was a finer corn cob grind. To obtain the finer corn cob
grind sample, the hammer milled corn cobs were further ground in a
0.25 kW (0.33 hp) Wiley mill (Thomas Scientific, Swedesboro, NJ)
fitted with 2.0 mm screen.

Table 1
Properties of corn cob grinds used for the study.

Grinding machine with Moisture content Particle size of Bulk density


the screen size used of the grind the grind (mm) of the grind
for grindinga (% w.b.) (n = 3) b
(n = 3) c
(kg m-3) (n = 3)b

Hammer mill with 7.5 ± 0.1 2.81 ± 0.31 229.0 ± 9.8


6.4 mm (1/4 in) screen
Wiley mill with 2.0 mm 7.4 ± 0.2 0.85 ± 0.25 316.3 ± 2.1
screen d
a
The screen size represents the diameter of the holes in the screen.
b
Mean ± standard deviation.
c
Geometric mean particle diameter (mm) ± geometric standard deviation.
d
Corn cobs hammer milled using 6.4 mm screen was re-ground in Wiley mill with
2.0 mm screen. Fig. 1. Schematic of piston-cylinder compression/densification apparatus.
N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565 561

uniaxial piston-cylinder compression (densification) apparatus and


the briquetting procedure involved are given in [26]. A brief
description of the briquetting procedure follows. An INSTRON
universal testing machine (Model 4206; Instron Corporation, Canton,
MA) was used to apply a mechanical pressure of 150 MPa for
briquetting the corn cob grind. The inner diameter and height of the
steel cylinder (i.e., die) were 18.8 mm and 300 mm, respectively. The
piston was made of brass and was 50 mm longer than the cylinder to
help eject briquettes from the cylinder. A separate apparatus was used
to allow ejection of briquettes from the cylinder after compression
(Fig. 2).
About 5.0 g of corn cob grind was added to the cylinder through a
funnel. Using a steel rod, the grind was stirred to help the flow of grind
from the funnel. For the cases with preheating to 85 °C, before
compression of corn cob grind, the grind along with the cylinder was Fig. 3. Modifications to the densification cylinder (die) during preheating of corn cob
heated using heating tapes to achieve a preheating temperature of grind.

85 °C (Fig. 3). During preheating, the top opening of the cylinder was
sealed with a specially designed vapor-tight cover. The outside of the
compression cylinder wall and bottom steel base of the cylinder were 25.4 mm min-1 (1.0 in. min-1) was applied to the piston to eject the
covered with heating tapes (120 V, 2.58 A, 310 W, and single phase; briquette from the cylinder. In this study, the cylinder was not
BH Thermal Corporation, Columbus, OH) to heat the grind along with lubricated during briquetting, but periodically cleaned using a
the cylinder. About 50 mm thick fiberglass insulation was used to vacuum cleaner.
cover the heating tapes to avoid heat loss. Temperature of the grind
was measured inside of the cylinder at the center and about 35.0 mm 2.3. Briquetting experiments
above the bottom of the cylinder (i.e., half the height of the grind
inside the cylinder). A hand-held thermocouple-temperature record- Corn cob briquettes were made for the following factorial combina-
er (DIGI-SENSE®, Model No. 91100-40; Cole-Parmer Instruments Co., tion of conditions: (i) one compression pressure, 150 MPa; (ii) two
Vernon Hills, IL) was used to monitor the temperature of the grind. particle sizes, a coarser corn cob grind obtained from the hammer mill
Immediately after reaching the grind temperature of 85 °C, the power fitted with a screen having 6.4 mm openings (i.e., geometric mean
to the heating tapes was turned off, the cylinder was mounted/bolted particle diameter of 2.81 mm± 0.31) and a finer corn cob grind ob-
to the base of the INSTRON, and the vapor-tight cover on the top of the tained from the Wiley mill fitted with a screen having 2.0 mm openings
cylinder was removed. Subsequently, the bottom of the piston was (i.e., geometric mean particle diameter of 0.85 mm± 0.25); (iii) two
inserted into the cylinder, and the top of the piston was connected to nominal grind moisture contents, 10% and 20% (w.b.); and (iv) two
the crosshead of the INSTRON by a pin connection. The piston temperatures, 25 °C (i.e., room temperature) and 85 °C (i.e., preheating
attached to the crosshead of the INSTRON was actuated to compress temperature). At each briquetting condition, ten briquettes (i.e., ten
the grind from zero pressure to a maximum pressure of 150 MPa at a replications) were made. In this study, no binders (additives) were used
constant (downward) speed of 25.4 mm min-1 (1.0 in. min-1) to make for briquetting.
briquettes. After the completion of compression, the piston was taken A preheating temperature of 85 °C was chosen because Kaliyan
out of the cylinder at a constant crosshead (upward) speed of and Morey [26] reported that the glass transition temperature of corn
25.4 mm min-1 (1.0 in. min-1). stover averaged 75 °C. To avoid moisture loss/migration and related
Next, the cylinder was removed from the base of the INSTRON and problems during densification in lab-scale apparatuses, preheating
placed on the briquette ejection apparatus, which was bolted to the temperature should be kept below 100 °C [26]. Furthermore, Kaliyan
base of the INSTRON (Fig. 2). A constant crosshead speed of et al. [21] reported that temperatures of around 85 °C can occur in
conventional ring-die pelleting machines due to frictional heating,
and thus, steam conditioning may be avoided.

2.4. Briquette properties

Immediately after ejection from the die, unit density of the


briquettes (i.e., density of individual briquettes) was measured. Then,
the briquettes were transferred to zip-lock plastic bags and stored for
one week at room temperature (about 25 °C). Unit density, durability,
percentage expansions in axial direction, radial direction and volume,
and moisture content of the briquettes were measured after one week
of storage.
The durability of briquettes is a measure of the ability of the
briquettes to withstand the destructive forces such as compression,
impact and shear during handling and transportation. The possible
amount of fines (dust) generated from the briquettes due to the
mechanical handling and transportation can be estimated by 100
minus the percent durability. Durability of briquettes was measured
according to ASABE Standard S269.4 [27]. Durability was calculated as
the percentage of the original mass of the briquettes retained on
16.0 mm screen after tumbling in a durability tester (Continental-
Agra Equipment, Inc., Newton, KS) at 50 rpm for 10 min. For each
Fig. 2. Schematic of briquette ejection apparatus in operation. replication of the durability test, five briquettes weighing about 25.0 g
562 N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565

were used. Only two replications were done for the durability
measurement because of lack of samples. Unit density of briquettes
was calculated from the mass, diameter, and height of the briquettes
[27]. Axial, radial, and volume expansions of briquettes were
calculated as the percentage increase in height, diameter, and volume,
respectively. Moisture content of the briquettes was measured based
on ASABE Standard S358.2 [23].

2.5. Statistical analyses

All statistical analyses were performed using PASW Statistics 17.0


software (SPSS, Inc., Chicago, IL) at 5% significance level. Paired-
samples t-tests were performed on densities of briquettes measured
immediately after ejection from the die and after one week of storage
at room temperature to determine if there were any significant
differences between the means of the briquette densities measured at
these two times. The unit density and durability of briquettes
measured after one week of storage at room temperature were used
for the rest of the statistical analyses. Tests on the analysis of variance
(ANOVA) using a univariate general linear model were performed on
the density and durability of briquettes to understand the main effects
rather than the interaction effects of the densification process
variables. Independent sample t-tests were conducted to investigate
if there were any significant differences between the means of
briquette densities and durabilities of any two experimental condi-
tions such as particle sizes of 0.85 mm versus 2.81 mm, or moisture
contents of 10% versus 20%, or preheating temperatures of 25 °C
versus 85 °C.

3. Results and discussion

3.1. Characteristics of corn cobs

Fig. 4 shows the pictures of whole corn cobs (Fig. 4A), and corn cob
grinds (Figs. 4B, C) used for this study. In this study, the bulk density of
whole corn cobs obtained from a seed corn variety was measured at
163.9 ± 3.0 kg m-3 (10.2 ± 0.2 lb ft-3) at 10.4% (w.b.) moisture con-
tent. Dunning et al. [17] reported that the bulk density of whole corn
cobs was about 210 kg m-3 (13 lb ft-3). Foley [2] documented that
ground corn cobs had bulk densities of 248 to 272 kg m-3 (15.5 to
17.0 lb ft-3). Table 1 provides the particle size and bulk density of the
corn cob grinds used for making briquettes in this study. Corn cob
grinds had bulk densities of 229.0 and 316.3 kg m-3 (14.3 and
19.8 lb ft-3) for the geometric mean particle diameters of 2.81 mm
and 0.85 mm, respectively (Table 1).
Table 2 compares the chemical compositions of the corn cobs, corn
stover, and switchgrass. The constituents such as lignin, protein,
starch, fat, and water soluble carbohydrates are potential “natural
binders” in the biomass materials [22]. These natural binders can be Fig. 4. Pictures of whole corn cobs (A) and corn cob grinds (B and C).
activated (softened or melted locally) either by high moisture or heat
or steam to use their binding functionality. Lignin and hemicellulose
were found to be amorphous thermoplastic materials which would
undergo plastic deformation at low pressures for temperatures in the According to Foley [2], the physical structure of the corn cobs is
range of their glass transition temperatures (i.e., softening tempera- composed of inner pith to the outer fine chaffy ring with the following
tures) [28]. Plastic deformation of particles would help make the average contents: 1.9% pith, 60.3% woody ring, 33.7% coarse chaff, and
particle–particle bonding permanent. In addition, during briquetting 4.1% fine chaff (“as is” basis). The woody ring of the corn cobs could be
or pelleting at temperatures close to the glass transition range, the a difficult component to process while grinding and densifying. Foley
natural binding components can be squeezed out of the particles, [2] also reported that the compositions of corn cobs were 45.6%
which can then make solid bridges between particles. After cooling, cellulose, 39.8% hemicellulose, 6.8% lignin, 0.02% starch, 2.8% protein,
these solid bridges will be hardened (i.e., curing process). This would 0.6% fat, 3.5% pectin, and 1.7% ash on a dry mass basis. Thus, the
make the briquettes or pellets strong and durable [21,29–31]. The ash composition data measured for the corn cobs used in this study
contents (i.e., mineral contents) in the biomass materials show their (Table 2) are similar to the composition data reported by Foley [2].
relative abrasiveness to densification equipment when there is high The ash content of the corn cobs used in this study was 1.8% (d.b.),
friction/shear during densification such as in pelleting. The higher the which is in the range of values given by Foley [2] and Morey and
mineral content, the higher the abrasion [22]. Thimsen [3].
N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565 563

Table 3
Properties of corn cob briquettes.

Briquetting conditions Properties of briquettes

Maximum pressure Preheating Grind moisture content Immediately after ejection from the die After one week of storage at room temperature
(MPa) temperature (°C) (% w.b.) (n = 3)a,b
Unit briquette density (kg m-3) (n = 5)a Unit briquette density (kg m-3) (n = 5)a Durability (%) (n = 2)a

Corn cob grind from hammer mill with screen size of 6.4 mm (particle size = 2.81 mm ± 0.31c)
150 25 9.4 ± 0.1 1030.7 ± 8.2 939.3 ± 8.9 0.0 ± 0.0
150 85 9.4 ± 0.1 1096.3 ± 15.0 1100.1 ± 16.2 92.3 ± 0.3
150 25 18.5 ± 0.1 744.8 ± 11.6 603.8 ± 19.1 0.0 ± 0.0
150 85 18.5 ± 0.1 817.8 ± 12.9 791.7 ± 37.5 0.0 ± 0.0d

Corn cob grind from Wiley mill with screen size of 2.0 mm (particle size = 0.85 mm ± 0.25c)
150 25 9.6 ± 0.2 1060.2 ± 7.4 971.2 ± 5.2 0.0 ± 0.0
150 85 9.6 ± 0.2 1112.1 ± 3.4 1120.3 ± 6.1 88.2 ± 0.9
150 25 19.1 ± 0.1 747.2 ± 15.9 651.0 ± 19.6 0.0 ± 0.0
150 85 19.1 ± 0.1 797.8 ± 9.5 776.7 ± 12.2 0.0 ± 0.0d
a
Mean ± standard deviation.
b
Moisture content of the grind before densification and preheating (if applicable).
c
Geometric mean particle diameter ± geometric standard deviation (n = 3).
d
Preheating of corn cob grinds to 85 °C at the nominal grind moisture content of 20% (w.b.) appeared to help bond the corn cob particles; however, some of the bonding sites were
not strong enough in the briquette to avoid disintegration of briquettes during the durability test. This was observed from the broken pieces of briquettes during the durability test.
The broken pieces of briquettes obtained from the durability test had larger particle sizes than the original grind sizes used for making the briquettes due to the stronger (localized)
bonding between one or more adjacent particles. High compression pressure and high temperature may bring moisture from the inside of the particles to the surface of the particles
causing free moisture between particles, which, after evaporation of moisture, may have resulted in weaker bonding forces between particles.

3.2. Densification characteristics the briquettes ranged from 17.2 to 24.9 mm. The nominal grind
moisture content of 20% (w.b.) resulted in longer briquettes due to
The results of the paired-sample t-tests showed that the unit higher expansion of briquettes after ejection from the die. The relaxed
densities of briquettes measured immediately after ejection from the density of briquettes made at 25 °C ranged from 603.8 to 971.2 kg m-3
die were significantly higher (by 56.6 kg m-3) than the values (Table 3). The higher moisture content (for both particle sizes) or the
measured after one week of storage at room temperature (P < 0.05) larger particle size (for both moisture contents) resulted in signifi-
(Table 3). This may be due to the change in the briquette mass and/or cantly lower briquette density (P < 0.05) (Table 3). The finer particle
volume (possibly due to drying and elastic-spring-back/expansion of size (0.85 mm) resulted in lower expansion in briquette volume than
briquettes) during the one-week storage (Table 4). the coarser particle size (2.81 mm) at both the nominal grind
The ANOVA analyses were performed to understand the main moisture contents of 10% and 20% (w.b.) (Table 4).
effects of the densification process variables on the density and For both the nominal grind moisture contents of 10% and 20% (w.
durability of corn cob briquettes. The ANOVA analyses revealed that b.), durability of corn cob briquettes made at 25 °C was 0% (Table 3).
both the density and durability of briquettes were significantly Kaliyan [31] reported that at nominal grind moisture contents of 10%
affected by particle size, moisture content, and preheating tempera- and 20% (w.b.), the briquette durabilities for corn stover were about
ture (P < 0.05). 75 and 96%, respectively. For switchgrass, the briquette durability was
0% at both 10% and 20% (w.b.) grind moisture contents [31]. The water
3.2.1. Densification without preheating soluble carbohydrates in the biomass may act as a binder in the
For densification of corn cobs at 25 °C, the diameter of the presence of enough moisture. Corn cobs and switchgrass have
briquettes produced ranged from 19.5 to 20.2 mm, and the length of relatively low levels of water soluble carbohydrates (1.1% d.b. for

Table 4
Moisture content and expansion of corn cob briquettes measured after one week of storage at room temperature.

Briquetting conditions Briquette moisture Mean expansion of briquettes (%) (n = 5)


content (% w.b.) (n = 3)a
Maximum pressure Preheating Grind moisture content Axial directionc Radial directiond Briquette volume
(MPa) temperature (°C) (% w.b.) (n = 3)a, b

Corn cob grind from hammer mill with screen size of 6.4 mm (particle size = 2.81 mm ± 0.31e)
150 25 9.4 ± 0.1 8.6 ± 0.1 6.4 1.1 8.6
150 85 9.4 ± 0.1 8.3 ± 0.1 -1.0 0.1 -0.8
150 25 18.5 ± 0.1 15.7 ± 0.4 14.4 2.3 19.6
150 85 18.5 ± 0.1 14.2 ± 0.2 0.01 -0.2 -0.3

Corn cob grind from hammer mill with screen size of 2.0 mm (particle size = 0.85 mm ± 0.25e)
150 25 9.6 ± 0.2 9.0 ± 0.2 6.0 1.1 8.3
150 85 9.6 ± 0.2 8.5 ± 0.2 -0.7 -0.3 -1.2
150 25 19.1 ± 0.1 15.5 ± 0.3 9.1 0.6 10.4
150 85 19.1 ± 0.1 14.2 ± 0.3 -1.2 -0.2 -1.5
a
Mean ± standard deviation.
b
Moisture content of the grind before densification and preheating (if applicable).
c
Negative values indicate the decrease in briquette length. This may have been due to the shrinkage of briquettes due to the moisture loss during storage for the nominal grind
moisture content of 10% (w.b.), and due to the loss of particles from the bottom of the briquettes due to the inadequate bonding of bottom layers of particles for the nominal grind
moisture content of 20% (w.b.).
d
Negative values indicate the decrease in briquette diameter. This may have been due to the shrinkage of briquettes due to moisture loss during storage.
e
Geometric mean particle diameter ± geometric standard deviation (n = 3).
564 N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565

Fig. 5. Corn cob briquettes made using coarser corn cob particles with geometric mean particle diameter of 2.81 mm ± 0.31 (pressure = 150 MPa, grind moisture content = 9.4% (w.b.),
and preheating temperature = 85 °C).

corn cobs and 2.2% d.b. for switchgrass) which may be insufficient to With preheating to 85 °C, the higher nominal grind moisture
make durable briquettes [21,31]. This may explain the zero percent content of 20% (w.b.) did not produce durable corn cob briquettes
briquette durability for corn cobs and switchgrass under conditions (Table 3). With preheating to 85 °C, at the nominal grind moisture
without preheating. But, corn stover has a higher level of water content of 10% (w.b.), durability of corn cob briquettes was
soluble carbohydrates (7.9% d.b.) which may have helped produce significantly higher for the coarser particle size (i.e., durability of
durable briquettes under conditions without preheating [21,31]. 92.3%) than for the finer particle size (i.e., durability of 88.2%)
(P < 0.05). This result suggests that the coarser particle size (geomet-
3.2.2. Densification with preheating ric mean particle diameter of 2.81 mm) may be preferable for
Figs. 5 and 6 present pictures of corn cob briquettes made using commercial production of corn cob briquettes/pellets. Using the
preheating to 85 °C. For densification of corn cobs at 85 °C, the coarser particle size could result in a lower cost for densification. The
diameter of the briquettes ranged from 19.0 to 19.1 mm, and the results also suggest that future work may be required to optimize the
length of the briquettes ranged from 15.6 to 20.8 mm. Compared to particle size, moisture content, and preheating temperature. Further-
the briquette densities at 25 °C, preheating to 85 °C resulted in more, the energy consumed for the lab-scale densification process is
significantly higher briquette densities at both 10% and 20% (w.b.) not likely to be an accurate estimate of a commercial scale operation.
nominal grind moisture contents for both the particle sizes tested Thus, future work is required to estimate the energy required for
(P < 0.05) (Table 3). The briquette densities of the preheated samples size reduction (hammer milling), preheating, and compressing in
were greater than 1100 kg m-3 at the nominal grind moisture content briquetting/pelleting machines at a pilot- or commercial-scale densi-
of 10% (w.b.) (Table 3). Preheating of corn cobs caused the briquette fication of corn cobs.
volume to shrink by 0.3 to 1.5%, possibly due to the loss of moisture
from the briquettes (Table 4).
At the nominal grind moisture content of 10% (w.b.), preheating to 4. Conclusions
85 °C produced corn cob briquettes with about 90% durability. It is
suggested that the improved binding of corn cob particles due to The following conclusions were drawn from this study.
preheating to 85 °C may have been due to the activation (softening) of
natural binders (i.e., lignin, protein, starch, and fat) that require heat • For briquetting at room temperature (about 25 °C), corn cob
for activation [31]. The activation/preheating temperature should be briquettes (diameter of about 20.0 mm) with relaxed densities of
at or above the glass transition temperature of the biomass materials 604 to 971 kg m-3 can be produced at a compression pressure of
in order to cause plastic deformation of particles and to fully activate 150 MPa, geometric mean particle diameters of 0.85 to 2.81 mm,
the natural binders [31]. The natural binder contents of corn cobs that and grind moisture contents of 10% to 20% (w.b.). However, the
require heat for activation are similar to those of the switchgrass durability of corn cob briquettes made at room temperature was
(Table 2). Kaliyan and Morey [26] reported that at a nominal grind zero percent.
moisture content of 10% (w.b.), switchgrass required preheating to at • For briquetting with preheating to 85 °C, corn cob briquettes
least 75 °C to produce durable briquettes (briquette durability of (diameter of about 19.0 mm) with relaxed densities of 1100 to
about 63%). Kaliyan and Morey [26] also found that at a nominal grind 1120 kg m-3 and durabilities of 88.2 to 92.3% can be produced at a
moisture content of 10% (w.b.), preheating to 75 °C produced highly compression pressure of 150 MPa, geometric mean particle diam-
durable corn stover briquettes (briquette durability of about 97%), eters of 0.85 to 2.81 mm, and grind moisture content of 10% (w.b.).
which may be due to the presence of high amounts of natural binders Preheating to 85 °C is necessary to activate the natural binders in the
that require heat for activation (Table 2). corn cobs such as lignin, starch, and protein to produce durable

Fig. 6. Corn cob briquettes made using finer corn cob particles with geometric mean particle diameter of 0.85 mm ± 0.25 (pressure = 150 MPa, grind moisture content = 9.6% (w.b.),
and preheating temperature = 85 °C).
N. Kaliyan, R.V. Morey / Fuel Processing Technology 91 (2010) 559–565 565

briquettes/pellets. The results are useful for making corn cob pellets [12] K.J. Shinners, G.C. Boettcher, J.T. Munk, M.F. Digman, R.E. Muck, P.J. Weimer,
Single-pass, split-stream of corn grain and stover: characteristic performance of
intended for commercial applications such as home heating.
three harvester configurations, ASABE Paper No. 061015, ASABE, St. Joseph, MI,
2006.
[13] S. Sokhansanj, A. Turhollow, J. Cushman, J. Cundiff, Engineering aspects of
Acknowledgements collecting corn stover for bioenergy, Biomass Bioenergy 23 (2002) 347–355.
[14] L.O. Pordesimo, W.C. Edens, S. Sokhansanj, Distribution of above-ground biomass
This article was presented at the 2008 American Society of in corn stover, Biomass Bioenergy 26 (2004) 337–343.
[15] W.H. Johnson, B.J. Lamp, Principles, Equipment and Systems for Corn Harvesting,
Agricultural and Biological Engineers (ASABE) Annual International Agricultural Consulting Associates, Inc., Wooster, OH, 1966.
Meeting in Providence, Rhode Island during June 29-July 2, 2008 as [16] R.D. Smith, J.B. Liljedahl, R.M. Peart, O.C. Doering, W.C. Sahm, Development and
ASABE Paper No. 084267. The authors thank the University of evaluation of a cob salvaging system, Trans. ASAE 27 (1984) 590–596.
[17] J.W. Dunning, P. Winter, D. Dallas, The storage of corncobs and other agricultural
Minnesota's Initiative for Renewable Energy and the Environment
residues for industrial use, Agric. Eng. 29 (1948) 11–13, 17.
(IREE) for providing support for this study, and Agricultural [18] R.D. Smith, R.M. Peart, J.B. Liljedahl, J.R. Barrett, O.C. Doering, Corncob property
Utilization Research Institute (AURI), Waseca, MN for generously changes during outside storage, Trans. ASAE 28 (1985) 937–942, 948.
supplying hammer milled corn cob samples. [19] J.M. Ebeling, B.M. Jenkins, Physical and chemical properties of biomass fuels,
Trans. ASAE 28 (1985) 898–902.
[20] K.J. Shinners, B.N. Binversie, P. Savoie, Harvest and storage of wet and dry corn
References stover as a biomass feedstock, ASAE Paper No. 036088, ASABE, St. Joseph, MI,
2003.
[1] R.C. Christiansen, Craving corn and the cob, Biomass Magazine, January 2009. [21] N. Kaliyan, R.V. Morey, M.D. White, A. Doering, Roll press briquetting and pelleting
Available at: http://www.biomassmagazine.com/article.jsp?article_id=2307. of corn stover and switchgrass, Trans. ASABE 52 (2009) 543–555.
[2] K.M. Foley, Physical Properties, Chemical Properties, and Uses of The Andersons’ [22] N. Kaliyan, R.V. Morey, Factors affecting strength and durability of densified
Corncob Products, 2nd ed. The Andersons, Cob Division Processing Group, biomass products, Biomass Bioenergy 33 (2009) 337–359.
Maumee, OH, 1978. [23] ASABE Standards, S358.2: Moisture measurement – Forages, ASABE, St. Joseph,
[3] R.V. Morey, D.P. Thimsen, Combustion of crop residues to dry corn, Proceedings of MI, 2003.
the ASAE Energy Symposium, Kansas City, Missouri, Sept. 29 - Oct. 1, 1980, 1980, [24] ASABE Standards, S319.3: Method of determining and expressing fineness of feed
p. 6. materials by sieving, ASABE, St. Joseph, MI, 2003.
[4] R.V. Morey, D.P. Thimsen, J.P. Lang, D.J. Hansen, A corncob–fueled drying system, [25] AOAC International, Official Methods of Analysis of AOAC International, 18th ed.
Trans. ASAE 27 (1984) 556–560. Association of Official Analytical Chemists, Gaithersburg, MD, 2005.
[5] H. Jiang, R.V. Morey, Air-gasification of corncobs at fluidization, Biomass [26] N. Kaliyan, R.V. Morey, Densification characteristics of corn stover and
Bioenergy 3 (1992) 87–92. switchgrass, Trans. ASABE 52 (2009) 907–920.
[6] J.M. Dominguez, N. Cao, C.S. Gong, G.T. Tsao, Dilute acid hemicellulose [27] ASABE Standards, S269.4: Cubes, pellets, and crumbles – Definitions and methods
hydrolysates from corn cobs for xylitol production by yeast, Bioresour.Technol. for determining density, durability, and moisture content, ASABE, St. Joseph, MI,
61 (1997) 85–90. 2003.
[7] W.T. Tsai, C.Y. Chang, S.L. Lee, A low cost adsorbent from agricultural waste corn [28] E.L. Back, N.L. Salmen, Glass transitions of wood components hold implications for
cob by zinc chloride activation, Bioresour. Technol. 64 (1998) 211–217. molding and pulping processes, Tappi 65 (1982) 107–110.
[8] F. Yu, R. Ruan, P. Chen, S. Deng, Y. Liu, X. Lin, Liquefaction of corn cobs with [29] H. Rumpf, The strength of granules and agglomerates, in: W.A. Knepper (Ed.),
supercritical water treatment, Trans. ASABE 50 (2007) 175–180. Agglomeration, John Wiley and Sons, New York, NY, 1962, pp. 379–418.
[9] C.L. Crofcheck, M.D. Montross, Effect of stover fraction on glucose production [30] L.G. Tabil Jr., Binding and Pelleting Characteristics of Alfalfa, Ph.D. dissertation,
using enzymatic hydrolysis, Trans. ASAE 47 (2004) 841–844. Department of Agricultural and Bioresource Engineering, University of Saskatchewan,
[10] J.J. Hanway, Growth stages of corn (Zea mays, L.), Agronomy J. 55 (1963) 487–492. Saskatoon, Saskatchewan, Canada, 1996.
[11] L.O. Pordesimo, B.R. Hames, S. Sokhansanj, W.C. Edens, Variation in corn stover [31] N. Kaliyan, Densification of Biomass, Ph.D. dissertation, Department of Biopro-
composition and energy content with crop maturity, Biomass Bioenergy 28 ducts and Biosystems Engineering, University of Minnesota, Twin Cities, MN,
(2005) 366–374. 2008.

You might also like