You are on page 1of 43

MALTODEXTRIN

Matthew J. Mollan Jr. and Metin Celik

Pharmaceutical Compaction Research Laboratory


College of Pharmacy
Rutgers, The State University of New Jersey
Piscataway, NJ 08855

ANALYTICAL PROFILES OF DRUG SUBSTANCES 307 Copyright 0 19% by Academic Ress, Inc.
AND EXCIF'IENTS-VOLUME 24 All rights of reproduction in any form reserved.
308 MATTHEW J. MOLLAN JR.AND METlN CELIK

CONTENTS

1. Description
1.1 Name, Definition, Formula
1.2 Appearance
1.3 Carbohydrate Profile
1.4 Uses and Applications

2. Physical Properties
2.1 Particle Morphology
2.2 Crystallographic Properties
2.3 Thermal Analysis
2.4 Particle Size Distribution
2.5 Surface Area
2.6 Mercury Poroshnetry
2.7 Density
2.8 Moisture
2.9 Powder Flow
2.10 Compaction
2.11 Viscosity

3. Methods of Analysis
3.1 Cornpendial Tests

4. Identification
4.1 Maltodextrin Saccharide Separations
4.2 Thin Layer Chromatography
4.3 Liquid Chromatography
4.4 Supercritical Fluid Chromatography
4.5 NMK Spectroscopy

5 . Acknowledgments

6. References
MALTODEXTRIN 309

1. Description
1.1 Name

Maltodextrins are also known as hydrolyzed cereal solids, and are starch
conversion products which contain a relatively small amount of dextrose
and maltose. The United States Pharmacopeia and National Formulary
[l] definition of a maltodextrin is: Maltodextrin is a non-sweet, nutritive
saccharide mixture of polymers that consist of D-glucose units, with a
Dextrose Equivalent less than 20. It is prepared by the partial hydrolysis
of a food grade starch with suitable acids andor enzymes. It may be
physically modified to improve its physical and functional
characteristics.

The chemical Abstracts identification number for Maltodextrin is AS-


9050-36-6.

Maltodextrin has a general formula of


H( C6H1005 >"-OH
and is composed of D-glucose units linked primarily by a-1-4 bonds.

Maltodextrin is listed as generally recognized as safe (GRAS)for human


consumption under 21CFR 184.1444.

Maltodextrins have both Food Chemical Codex and National Formulary


monographs.

Starch conversion products with a dextrose equivalent values above


twenty are referred to as corn syrup solids. Starch conversion products
with only a trace amount of dextrose are known as dextrins. Starch
conversion products having a dextrose equivalent value not substantially
above twenty, and containing small amounts of dextrose are known as
maltodextrins. Kanig [2] described the dextrose content of a
maltodextrin as less than about 2.4% by weight, and the amount of
maltose as less than about 9.0% by weight.

The term Dextrose Equivalent Value, D.E., is defined as the reducing


value of the hydrolysate material compared to the reducing value of an
equal weight of dextrose expressed as a percent, dry basis.
310 MATTHEW J. MOLLAN JR. AND METlN CELIK

Reducing Value of Hydrolysate Material


D.E. = x 100
Reducing Value of Dextrose

The dextrose equivalent value is an indicator of the degree of


depolymerization of starch, and the D.E. value will influence the
physical properties of the maltodextrin. The higher the D.E. value, then
the greater the extent of starch hydrolysis.

1.2 Appearance

Maltodextrins are white to off white powders or granules. Maltodextrins


are bland, odorless, with a low sweetness level. The materials are often
physically processed to improve their handling characteristics.

Maltodextrins are produced from starch, usually corn. The starch, which
is almost pure carbohydrate, is cooked or pasted to open the granule and
then hydrolyzed. Products can be made by hydrolyzing with acid or
enzymes or with a combination of acid and enzymes. After the desired
amount of hydrolysis has occurred, the reaction is stopped, the product is
filtered to remove insoluble materials, then dried.[3]. The average
molecular weight decreases as the dextrose equivalent value of the
maltodextrin increases, but even at low D.E. values, it is much smaller
than the original starch. This relative molecular weight difference
between starch and the hydrolyzed sugars gives the maltodextrins a
portion of their valuable functional properties for the food and
pharmaceutical industry.

1.3 Carbohydrate Profile

Maltodextrins will have different carbohydrate profiles depending on


their dextrose equivalent value. The carbohydrate profile of a
maltodextrin has important effects for the physicochemical properties of
the maltodextrin. For example, the low molecular weight components
will influence sweetness, viscosity, and humectant properties, while the
MALTODEXTRIN 31 1

high molecular weight components will influence solubility and solution


stability.

Maltrin@M150 [4] maltodextrin


Standard Specifications
Dextrose Equivalent 13.0 - 17.0
Carbohydrate Profile (Dry Basis)
Monosaccharides 0.7%
Disaccharides 4.5%
Trisaccharides 6.6%
Tetrasaccharides 5.3%
Pentasaccharides and Above 82.9%

Maltrin@M5 10 [4]maltodextrin
Standard Specifications
Dextrose Equivalent 9.0 - 12.0
Carbohydrate Profile (Dry Basis)
Monosaccharides 0.8%
Disaccharides 2.9%
Trisaccharides 4.4%
Tetrasaccharides 3.8%
Pentasaccharides and Above 88. I %

Maltodextrin [5]
Standard Specifications
Dextrose Equivalent 12.1
Carbohydrate Profile (Dry Basis)
Monosaccharides 0.9%
Disaccharides 2.5%
Trisaccharides 4.0%
Tetrasaccharides 3.4%
Pentasaccharides and Above 89.2%

1.4 Uses and Applications

Maltodextrin is generally used in chewable tablet formulations, however


the modified forms can be used in oral tablet formulations. It is also
used in tablet coating formulations. Some uses of maltodextrins which
312 MATTHEW J. MOLLAN JR. AND METIN CELIK

have been given patent protection are: US patent # 3873694 [2] for use in
direct compression tabletting; South African patent # ZA 800209 A [6]
for use in coating: and South African patent # ZA 5000209 A [7] for use
in coating.

'Tablet formulations containing a significant percentage of high D.E.


value maltodextrin must guard against moisture uptake at high
humidities since the material is hygroscopic. Tablets containing a
significant portion of maltodextrin have extended disintegration times in
water, (greater than five minutes), which may influence dissolution
behavior.

Maltodextrins are used extensively in the food industry as a moisture


conditioner, food plasticizer, crystallization inhibitor, stabilizer, carrier
and bulking agent [8]. The material is often used pharmaceutically in
solid dosage forms as a tablet fillerhinder excipient, but is usually
physically processed by spray drying, fluidized bed agglomeration, and
roller compaction to improve its physical characteristics.

Maltodextrins are generally considered not to be susceptible to undergo


the Maillard reaction, which leads to browning and discoloration. A
paper by Schmidt and Brogmann [9] did find discoloration of ascorbic
acidkodium bicarbonate effervescent tablets containing maltodextrin as a
tablet binder. The found the intensity of discoloration increased with the
degree of degradation of dextrins and maltodextrins, and was
proportional to their dextrose equivalent value.

Maltodextrins have been used as binding agents in wet granulation


processes [lo], and as direct compression binding agents [ l 1, 12, 13, 141
for tablet formulations. They have also been used as stabilizers in solid-
state emulsion systems [ 151.

2. Physical Properties

The physical properties of maltodextrins are determined by their


respective dextrose equivalent value, their saccharide profile, and by the
method by which they were physically processed for use.
MALTODEXTRIN 313

2.1 Particle Morphology

Scanning electron photomicrographs of some maltodextrins are shown in


Figures 1-5. The photomicrographsare at 1OOX, 400X, and 1200X
magnification, from top to bottom, respectively. Figure 1 is of a roller
compacted maltodextrin, (Experimental Maltodextrin, Edward Mendell
Co.), and shows a very dense structure with many small particles
adhering to the surface. Figure 2 is of a spray dried maltodextrin,
(Maltrin@M5 10, Grain Processing Co.), and shows a relatively smooth
surfaced material with a generally round shape. At higher
magnifications, large pores are distinctly seen. Figure 3 is of a fluidized
bed agglomerated maltodextrin, (Maltrin@M500, Grain Processing Co.),
and appears as a very porous mass with many irregularities. Figures 4
and 5 are of the fluidized bed agglomerated maltodextrins, (Malta*Gran
TG@,and Malta*Gran lo@,ZumbroAFP Inc.) and illustrate the smooth
surface and large pores which exist from the fluidized bed agglomeration
method.

2.2 Crystallographic Properties

The x-ray powder diffraction pattern of the maltodextrins which had


SEM photomicrographstaken are shown in Figures 6-10. The samples
all behaved similarly, and were considered amorphous materials
according to x-ray powder diffraction.

2.3 Thermal Methods of Analysis

Differential scanning calorimetry thermograms of maltodextrins did not


show any significant events such as glass transitions or melting
endotherms in the temperature range examined, 25OC to 225OC. A slight
endotherm was seen from approximately 6OoCto 110°C due to
desorption of unbound water [161.

Thermogravimetricweight loss for maltodextrins was found to have a


total volatile content with a range of 7.0%-7.4%, and the individual
FIGURE 1: Scanning Electron Photoinicrograph ot
Experimental Maltodextrin

314
FIGURE 2: Scanning Electron Photomicrograph of
Maltrin M510

315
FIGURE 3: Scanning Electron Photomicrograph of
Maltrin M500

316
FIGURE 4: Scanning Electron Photomicrograph of
Malta*Gran TG

317
FlGURE 5: Scanning Electron Photomicrograph of
Malta*Gran 10
MALTODEXTRIN 319

2.00,

i .ao.

1.60.

1.40'

1.20.

1.00-

0.80.

0.60

0.40'

0.20'

0.0 5.0 10.0 t5.0 20.0 25.0 30.0

FIGURE 6 : Powder X-Ray Diffraction Pattern of


Experimental Maltodextrin

1.80.

1.60'

1.40'
I

i .OO'

8p%
*w
0.80.

0.60' /Ifr
0.40-

/ / F -
0.20-

. ._,..- ,__ - 1 , - - I- , - . I

FIGURE 7: Powder X-Ray Diffraction Pattern of


Maltrin M510
320 MATTHEW J. MOLLAN JR. AND METIN CELlK

i.aa.

1 60-

0 a01
i
0 60i

FIGURE 8: Powder X-Ray Diffraction Pattern of


Maltrin M500

1 40

FIGURE 9: Powder X-Ray Diffraction Pattern of


Malta*Gran TG
MALTODEXTRIN 321

x io=
2.00

l._/

1.60'

1.40'

1.20-

0.0 5.0 10.0 15.0 20.0 25.0 90.0

FIGURE 10: Powder X-Ray Diffraction Pattern of


Malta*Gran 10
122 MATTHEW J. MOLLAN JR.AND METIN CELIK

results are shown in Table 1. All samples behaved similarly, and no


bound waters were discernable.

2.4 Particle Size Distribution

The particle size distributions of maltodextrins available vary


considerably, depending on the method of processing the material. Very
fine grades of maltodextrin powder are available, however the particle
sizes of the physically processed maltodextrins are much larger. The
particle size distribution, by sieve analysis, of several physically
processed maltodextrins are shown in Figure 1 1. The fluidized bed
agglomerated maltodextrins had the largest particle size because they are
created from spray dried maltodextrin which is then agglomerated with
water in a fluidized bed process.

2.5 Surface Area

The surface areas of the maltodextrins was recorded by a krypton gas 3-


point BET method [ 171. The surface area analysis results are shown in
Table 2. The spray dried and fluidized bed agglomerated maltodextrins
had similar surface areas, while the roller compacted maltodextrin had a
much larger surface area. These results were consistent with the
observations made by electron microscopy, which illustrated the smooth
surfaces of the spray dried and fluidized bed agglomerated maltodextrins,
while showing the rough surface of the roller compacted maltodextrin.

2.6 Mercury Porosimetry

The measurement of pore volume and distribution of maltodextrins was


performed by the method of Ritter and Drake [ 18). Figures 12 and 13
are from the results of the incremental amount of mercury intruded into
the maltodextrin powder vs pore diameter. Figure 12 shows the fluidized
bed agglomerated maltodextrins to have similar profiles, with
Malta*Gran 1Om having a larger number of pores due to its larger mean
particle size. The spray dried maltodextrin had a smaller mean pore size
than the fluidized bed agglomerated maltodextrins. Figure 13 is the
Total Volitile Content
Maltodextrin Processing Method
from TG (%)
Experimental
roller compaction 7.0
Maltodextrin
Maltrin" M 5 10 spray drying 7.2
fluidized bed
Maltrin" M500 7.4
agglomeration
fluidized bed
Malta'Gran TG" 7.2
agglomeration
fluidized bed
Malta 'Gran 10" 7.0
agglomeration

TABLE 1: Thermogravimetric Weight Loss of the Maltodextrins


323 M A T H E W J. MOLLAN JR. AND METIN CELIK

100

90
0 ,

ir!
30
e
c4 20
10

0 I I ' / ' I ' I ' I I / ' l '

0 100 200 300 400 500 600 700 800

PARTICLE SIZE (microns)

FIGURE 11: Particle Size Distribution of the Maltodextrins


0 = Experimental Maltodextrin
= Maltrin M510
0 = MaltrinM500
r = Malta*GranTG
V = Malta*Gran 10
Surface Area
Maltodextrin Processing Method
(m2/g)
Experimenta I
roller compaction 1.66
Maltodextrin
Maltrin" M 5 1 0 spray drying 0.12
fluidized bed
Maltrin" M 5 0 0 0.12
agglomeration
fluidized bed
Malta*Gran TG" 0.10
agglomeration
fluidized bed
Malta'Gran 10" 0.14
agglomeration

TABLE 2: Surf;ice Area of the Maltodextrins


326 MATTHEW J. MOLLAN JR. AND METIN CELIK

1
1 0.0

Pore Diameter (um)

FIGURE 12: Mercury Porosimetry


0 = Experimental Maltodextrin
0 = Maltrin M510
0 = Maltrin M500
V = Malta*Gran 10
MALTODEXTRIN 321

T
t" 0.
a
1
V
0
1
U
m
0.
e

0.
.o
Pore Diameter (um)

FIGURE 13: Mercury Porosimetry: log log scale


0 = Experimental Maltodextrin
0= Maltrin M510
0 = Maitrin M500
V = Malta*Gran 10
328 MATTHEW J. MOLLAN JR.AND METIN CELIK

same data plotted on a log-log scale and illustrates the roller compacted
maltodextrins large number of small pores, which accounts for its large
surface area as compared with spray dried or fluidized bed agglomerated
maltodextrins.

2.7 Density

The densities of the maltodextrins are shown in Table 3 . True density


was determined by helium pycnometry, while tap density (100 taps) and
bulk density were determined in a 100 mL glass cylinder. The densities
of the maltodextrins differ due to their method of processing. This is
especially shown in the "true density" measurements, which does not
measure internal or ''closed" pores, only external or open pores.

2.8 Moisture

The moisture content of the maltodextrins depends on the relative


humidity of the surrounding atmosphere, and maltodextrins can sorb and
desorb significant amounts of water. Figure 14 illustrates the sorption
isotherm of the maltodextrins examined. They exhibited very similar
profiles to each other despite differences in their method of processing.
The maltodextrin powders began to "gel", and formed a pliable mass
when stored at conditions above 75% relative humidity. The United
States Pharmacopeia and National Formulary states that maltodextrin
should be stored in tight containers or well closed containers at a
temperature not exceeding 3OoC and a relative humidity not exceeding
50% [ 13. Hygroscopicity and D.E. value generally correlate, with the
degree of hygroscopicity increasing with the increase in D.E. value.

2.9 Powder Flow

Flow parameters measured for the maltodextrins included: flow through


and orifice, angle of repose, and percent compressibility [19]. The
results are shown in Table 4. The maltodextrins tested were all free
flowing, and the gravimetric method of describing powder flow showed
the greatest difference between the samples due to density
Bulk Density Tap Density True Density
Maltodextrin
(g/ml) (g/mll (g/ml)
Experimental Maltodextrin .57 .65 1.503
Maltrin M510 .50 .56 1.425
Maltrin M500 .27 .32 1.410
I I I

Malta *Gran TG .38 .44 1.424


Malta *Gran 10 .29 .35 1.417

TABLE 3: Densities of the Maltodextrins


330 MATTHEW J. MOLLAN JR. AND METIN CELIK

50 -
45 -

0 10 20 30 40 50 60 70 80 90 100

ZELATIVE HUMIDITY ( X i

FIGURE 14: Sorption Isotherm of the Maltodextrins


0 = Experimental Maltodextrin
= Maltrin M510
0 = Maltrin M500
v = Malta*Gran TG
V = Malta*Gran 10
Maltodextrin

TABLE 4: Flow Paraineters of the Maltodextrins


332 MATTHEW J. MOLLAN JR. AND METIN CELIK

considerations, i.e. high density materials generally tend to flow better


than low density materials. Other very fine grade maltodextrin powders
often will not flow, and this one of the primary reasons for physically
processing maltodextrins for pharmaceutical use.

2.10 Compaction

Compression of a maltodextrin powder mass in a die causes the


formation of a compact. Compaction of maltodextrin powder, (with
0.5% magnesium stearate added as a lubricant) into a tablet was
performed at several pressures and the results are shown in Figure 15.
The maltodextrins form relatively strong tablets at modest compressional
forces, and can be considered to have good binding ability.

2.1 1 Viscosity

Maltodextrins exhibit Newtonian viscosity, and decrease in viscosity as


they are heated. They generally exhibit low viscosity at modest solids
content. Kenyon and Anderson [3] presented data of viscosity of
maltodextrin solid solutions at varied percent solids. They stated that
maltodextrins will demonstrate good solubility in the following solids
range: 5 D.E. = 30-45%; 10 D.E. = 45-55%; 15 D.E. = 5045%; and 20
D.E. = 60-75%.

3. Methods of Analysis

3.1 Compendia1 Tests

According to the United States Pharmacopeia and National Formulary


[ 11, maltodextrin is tested according to/for: microbial limits, pH, loss on
drying, residue on ignition, heavy metals, protein, sulfur dioxide, and
dextrose equivalent.
M ALTODEXTRIN 333

275

250

w 175
0
d
0 150
Lr,
0 125
i5
z 100

2
CA
75
50

25

0
0 50 100 150 200 250 300 350 400 450

APPLIED PRESSURE (ma)

FIGURE 15: Tablet Crushing Force vs Applied Pressure of the


Maltodextrins
0 = Experimental Maltodextrin
0 = Maltrin M510
0 = Maltrin M500
v = Malta*GranTG
V = Malta*Gran 10
334 MAlTHEW J. MOLLAN JR. AND METIN CELIK

. . .
Microbial hmlts. b e t h o d <6 1 >]
It meets the requirements of the tests for absence of Salmonella species
and Escherichia coli.

g& [method <79 1 >]


The pH is between 4.0 and 7.0, in a 1 in 5 solution in carbon dioxide-free
water.

Loss ondrv- i n5. lmethod <73 I>]


Dry at 105°C for 2 hours in a forced-air oven: it loses not more than
6.0%of its weight.

Residue on ignition: [method <281>]


The residue on ignition is not more than 0.5%.

Heavy Metals: [method I1 <23


Not more than 5 ppm heavy metals can be present

Protein:
Transfer about 10 g of Maltodextrin, accurately weighed, to an 800-mL
Kjeldahl flask, and add 10 g of anhydrous potassium sulfate or sodium
sulfate, 300 mg of copper selenite or mercuric oxide, and 60 mL of
sulfuric acid. Gently heat the mixture, keeping the flask inclined at about
a 45' angle, and after frothing has ceased, boil briskly until the solution
has remained clear for about I hour. Cool, add 30 mL of water, mix, and
cool again. Cautiously pour about 75 mL (or enough to make the
mixture strongly alkaline) of sodium hydroxide solution (2 in 5) down
the inside of the flask so that it forms a layer under the acid solution, and
then add a few pieces of granular zinc. Immediately connect the flask to
a distillation apparatus consisting of a Kjeldahl connecting bulb and a
condenser, the delivery tube of which extends well beneath the surface of
an accurately measured excess of 0.1 N sulfuric acid contained in a 50-
mL flask. Gently rotate the contents of the Kjeldahl flask to mix, and
distill until all ammonia has passed into the absorbing acid solution
(about 250 mL of distillate). To the receiving flask add 0.25 mL of
methyl red-methylene blue TS, and titrate the excess acid with 0.1 N
sodium hydroxide. Perform a blank determination, substituting pur.:
sucrose or dextrose for the test specimen, and make any necessary
correction. Each mL of 0.1 N sulfuric acid consumed is equivalent to
MALTODEXTRIN 335

1.401 mg of nitrogen (N). Calculate the percentage of N in the specimen


taken, and then calculate the percentage of protein by multiplying the
percentage of N by 6.25. The limit is 0.1%.

Sulfur dioxide
Hydrogen peroxide solution:
Dilute 30 percent hydrogen peroxide with water to obtain a 3% solution.
Just before use, add 3 drops of methyl red TS, and neutralize to a yellow
endpoint with 0.01 N sodium hydroxide. Do not exceed the endpoint.

Nitrogen:
Use high-purity nitrogen, with a flow regulator that will maintain a flow
of 200 f 10 mL per minute. Guard against the presence of oxygen by
passing the nitrogen through a scrubber, such as alkaline pyrogallol,
prepared as follows. Add 4.5 g of pyrogallol to a gas-washing bottle,
purge the bottle with nitrogen for 3 minutes, and add a solution
containing 85 mL of water and 65 g of potassium hydroxide, while
maintaining an atmosphere of nitrogen in the bottle.

Apparatus:
The apparatus (see Figure I ) is designed to effect the selective transfer of
sulfur dioxide from the specimen in boiling aqueous hydrochloric acid to
the Hydrogen peroxide solution. The backpressure is limited to the
unavoidable pressure due to the height of the Hydrogen peroxide solution
above the tip of the bubbler, F. Keeping the backpressure as low as
possible reduces the likelihood that sulfur dioxide will be lost through
leaks. Preboil vinyl and silicone tubing. Apply a thin film of stopcock
grease to the sealing surfaces of all the joints except the joint between the
separatory funnel and the flask, and clamp the joints to ensure tightness.
The separatory funnel, B, has a capacity of 100 mL or greater. The inlet
adapter, A, with a hose connector provides a means of applying
headpressure over the solution. [Note: A pressure-equalizing dropping
funnel is not recommended because condensate, which may contain
sulfur dioxide, is deposited in the funnel and the side arm.] The round-
boom flask, C, is a 1000-mL flask with three 24/40 tapered joints. The
gas inlet tube, D, is long enough to permit introduction of the nitrogen
within 2.5 cm of the bottom of the flask. The Allihn condenser, E, has a
jacket length of 300 mm. The bubbler, F, (see Figure 11) is fabricated
336 MATTHEW J. MOLLAN JR. AND METIN CELIK

FIGURE I. Apparatus for Sulfur Dioxide Test


(reproduced with permission of USP)
MALTODEXTRIN 331

from glass according to the dimensions given in Figure 11. The


Hydrogen peroxide solution is contained in a vessel, G, having an inside
diameter of about 2.5 cm and a depth of about 18 cm. Circulate coolant,
such as a mixture of water and methanol (4:1) maintained at 5', to chill
the condenser.

Procedure:
Position the Apparatus in a heating mantle controlled by a power-
regulating device. Add 400 mL of water to the flask. Close the stopcock
of the separatory funnel, and add 90 mL of 4 N hydrochloric acid o the
separatory funnel. Begin the flow of nitrogen at a rate of 200 f 10 mL
per minute. Start the condenser coolant flow. Add 30 mL of Hydrogen
peroxide solution to vessel, G. After 15 minutes, remove the separatory
funnel, and transfer a mixture of 50.0 g of Maltodextrin, accurately
weighed, and 100 mL. of alcohol solution (5 in 100). Apply stopcock
grease to the outer joint of the separatory funnel, return the separatory
funnel to the tapered joint flask, and concomitantly resume the nitrogen
flow. Apply headpressure above the hydrochloric acid solution in the
separatory funnel with a rubber bulb equipped with a valve. Open the
stopcock of the separatory b e 1 to permit the hydrochloric acid solution
to flow into the flask. Continue to maintain sufficient pressure above the
hydrochloric acid solution to force it into the flask. [Note: The stopcock
may be temporarily closed, if necessary, to pump up the pressure.] to
guard against the escape of sulfur dioxide into the separatory funnel,
close the stopcock before the last few mL of hydrochloric acid drain out.
Apply power to the heating mantle sufficient to cause about 85 drops of
reflux per minute. After refluxing for 1.75 hours, remove vessel G, add
3 drops of methyl red TS, and titrate the contents with 0.01 N sodium
hydroxide VS, using a 10-mL burette with an overflow tube and a hose
connection to a carbon dioxide-absorbing tube, to a yellow endpoint that
persists for a least 20 seconds. Perform a blank determination, and make
any necessary correction (see Titrimetry <23 l>). Calculate the quantity,
in pg, of SO, in each g of the Maltodextrin taken by the formula:

in which 32.03 is the milliequivalent weight of sulfur dioxide, V is the


volume, in mL, of titrant consumed, N is the normality of the titrant, and
338 MATTHEW J. MOLLAN JR.AND METIN CELIK

FIGURE 11. Bubbler (F) for Sulfur Dioxide Apparatus


(reproduced with permission of USP)
MALTODEXTRIN 339

W is the weight, in g, of Maltodextrin taken. Not more than 40 pg of


sulfur dioxide per gram of Maltodextrin can be found (.004%).

Dextrose Eauivalent;

Standard solution:
Dissolve an accurately weighed quantity of USP Dextrose RS in water,
and dilute quantitatively with water to obtain a solution having a known
concentration of about 10 mg per mL.

Test solution:
Transfer about 5 g of Maltodextrin, accurately weighed, with the aid of
hot water to a 100-mL volumetric flask, cool, add water to volume, and
mix.

Procedure:
Transfer 25.0-mL portions of alkaline cupric tartrate TS to each of two
boiling flasks. Bring the contents of one flask to boiling within about 2
minutes while titrating with Standard solution to within 0.5 mL of the
anticipated endpoint. Boil gently for 2 minutes. Continue to boil gently,
add 2 drops of methylene blue solution (1 in loo), and complete the
titration within 1 minute by adding the Standard solution dropwise or in
small increments until the blue color disappears, determined by viewing
against a white background in daylight or under equivalent illumination.
If more than 0.5 mL of the titrant was required after the addition of the
indicator, repeat the titration, adding the necessary volume of titrant
before adding the indicator. Bring the contents of the second flask to
boiling, and similarly titrate with Test solution. Calculate the Dextrose
equivalent, on the dried basis, by the formula:

[ 1OO/( 1 - 0.01A)](C,/C,)(V,N,),

in which A is the percentage Loss on drying of the Maltodextrin taken,


C, is the concentration, in mg per mL, of Maltodextrin in the Test
solution, C, is the concentration, in mg per mL, of USP Dextrose RS in
the Standard solution, and V, and V, are the titrated volumes, in mL, of
the Test solution and the Standard solution, respectively. The Dextrose
equivalent is not more than 20. [Note: This is a limit test. For
340 MATTHEW J. MOLLAN JR. AND METIN CELIK

maltodextrins with lower reducing values, other procedures may give


other results.]

4. Identification

41 Maltodextrin Saccharide Separations

Maltodextrins can be considered non to weakly UV absorbing


compounds.

Maltodextrins are often characterized by their D.E. values, yet the


physicochemical properties of the maltodextrins are dependent on their
overall saccharide profile present in the final hydrolysate. Various liquid
chromatographic separation techniques for the separation of
oligosaccharides, including maltodextrins, have been recently reviewed
by Chums [20]. Typical phase systems applied are: (a) a reversed
acetonitrile + water gradient on aminopropyl- or cyanopropyl-modified
silica; (b) an aqueous mobile phase on octadecyl-modified silica; (c) an
acidic (PH 5 2) aqueous mobile phase on microparticulate cation-
exchange resins such as an Aminex HPX-22H phase [21]; and (d) anion
exchange chromatography using an alkaline (PH 2 13) aqueous mobile
phase and a sodium acetate gradient [22]. The separation of saccharides
on the basis of chain length ( DP = degree of polymerization) in which
DP 1 is glucose, DP2 is disaccharides, etc., becomes increasingly difficult
above DPlO because of the extremely low levels which are found in
maltodextrins.

4.2 Thin Layer Chromatography (TLC)

Classical TLC is of limited use in attempting to separate and identify


simple saccharides in maltodextrins. Covacevich and Richards in 1976
[23J used thin layer chromatography in the continuous mode at 30 2" *
using 20 x 20 cm precoated plates. The solvent used was n-propanol-
nitromethane-water (50:20:30). The compounds were detected by
spraying the plate with 50% sulfuric acid followed by heating at 110°C
for 30 minutes. They were able to separate isomaltodextrins up to DP9.
MALTODEXTRIN 341

Bosch-Reig et al. [24] used monodimensional TLC to separate a aqueous


solution of maltodextrins in several biological fluids. They used
cellulose plates with two different eluents: (a) n-butanoYethanol/water
(3:2:2), and (b) pyridine/ethyl acetate/acetic acidwater (5:5: 1:3). The
maltodextrins were located with silver nitrate reagent. They were able to
separate maltodextrins from DP2 to DP7 components. More recently,
the use of high performance thin layer chromatography,HPTLC, has
been used successfully in the analysis of maltodextrins. Vajda and Pick
[25] used HPTLC to separate maltodextrin hydrolyzate on HPTLC silica
plate. They used an eluent of 60% acetonitrile: 10% isopropanol; 15%
aqueous ammonia 0.3%: 15% aqueous potassium chloride 0.67M, with a
membrane pressure of 5 bars, flow rate of 0.04 mL/min, and a
development distance of 170 mm. They separated up to DP14, but were
unable to separate glucose and fructose.

4.3 Liquid Chromatography (LC)

High resolution techniques are essential when performing carbohydrate


analysis, since carbohydrates have a number of isomers and homologs
that structurally resemble one another. Since maltodextrins do not
exhibit characteristic absorbances in useful regions of the UV spectrum,
derivatization is often used. Post column labeling has the advantage of
direct injection of intact samples, but has the limitation of sensitivity of
detection due to poor yields of derivatives. Precolumn labeling is often
used because high yields of derivatives can be easily achieved. If an
m i n e type of column is used in HPLC analysis of maltodextrins, then
DPl is the saccharide which will elute first. Increases in chain length
then cause increases in elution time. If a resin based column is used,
then the high molecular weight material, high DP, will elute initially and
the DP1 will elute last.

Scobell, Brobst, and Steele [26] described an automated liquid


chromatographic system for quantitative analysis of carbohydrate
mixtures. In 1981, Scobell and Brobst [27] discussed problems in
separating oligosaccharides, and specifically when using cation-exchange
resins that the loss of resolution for higher oligosaccharides is probably
due to the unique helical structure of a-1-4 linked corn-derived
oligosaccharides. They also used clean-up procedures prior to analysis,
342 MATTHEW J. MOLLAN JR. AND METIN CELIK

since salts, acids, soluble proteins, and particulate matter will interfere
with the chromatographic analysis. They found that the use of silver
form resins provided superior analysis to that of equivalent resins in the
calcium form.

Cheetham, Sirimanne, and Day 2281 used a C,,-bonded silica column


with water as an eluent at a flow rate of 1 .O mL/min. They were able to
separate maltodextrins to a DP6, although pairing was seen above DP3.
The pairs of peaks was attributed to the a and b anomers of the
oligosaccharide. They then performed a sodium borohydride reduction
which replaced each pair of peaks with a single peak. This peak was
then taken to be due to the corresponding oligosaccharide alditol.

Warthesen [29] separated maltodextrins by an HPLC column (HPX-42A


from Bio-Rad laboratories) with two precolumns, cation exchanging and
anion exchanging, and used water as the mobile phase. Detection was by
differential refractive index (RI) detector. With a flow rate of 0.5
ml/min, the chromatographic analysis was completed for a maltodextrin
in 21 minutes and gave resolution of DPl to DPlO. The lowest
detectable level for each saccharide was about 2 pg injected or 0.1%
expressed as a dry weight basis. He overcame the problems of
insolubility of some carbohydrate material, and nonlinearity of the large
molecular weight peak, by using external standards for quantitation
instead of area normalization.

Brooks and Griffin 1301 examined corn syrup solids and maltodextrins
and characterized the water soluble saccharides. The DP 1 - 10 saccharide
components were separated by using a plastic cartridge C18 Resolve
column compressed in a radial compression module (Waters Associates)
with water as the mobile phase and detection by a differential
refractometer. Pairs of peaks were obtained for most components
between DP3 and DPIO. The peaks were attributed to the a and p
anomers of the saccharides. An external glucose standard was used to
obtain percent composition. They also determined overall molecular
weight profiles by High Performance Size Exclusion Chromatography
(HPSEC) with E-HighA and E-500 pBondage1 silica gel permeation
columns (Waters Associates) connected in series. Water containing
0.15M NaCl was used as the mobile phase. The HPSEC data showed
MALTODEXTRIN 343

that as the D.E. of the hydrolysate increased, the amount of soluble high
molecular weight saccharides decreased.

McGinnis, Prince and Lowrimore [313 used reverse-phase HPLC with a


refractive index detector. A reverse phase column was used since it has
wide availability, high capacity, and can use water as an eluent.
Adjustment of the polarity of the carbohydrate was needed since their is
very little interaction between the carbohydrate and the column. They
found the best column for separation of a mixture of maltodextrins was
with a ODs-2 packed with CIS(Whatman c-18). The retention times on
the column depended on the type of sugar unit, linkage, the anomeric
configuration, and the molecular weight.

Honda et al. [32] recently found that 1-phenyl-3-methyl-5-pyrazolone


reacts with reducing carbohydrates almost quantitatively under mild
conditions to yield strongly UV-absorbing and electrochemically
sensitive derivatives. A homologous series of isomaltodextrins was
separated with precolumn labeling.

Niessen et al. [33] in 1992 used on-line liquid chromatography/mass


spectrometry (LC/MS) for the analysis of maltodextrins, and could detect
oligomers up to a DP 10. Their methodology used a mobile phase of 50
mmol/L ammonium acetate, with gradient elution with 0-50% methanol
and the use of a octadecyl column. Peak splitting occurred at higher DP
values due to separation of the two anomeric forms of the sugar.
LC/thermosprayMS analysis of maltodextrin was performed with 1O4 M
aqueous sodium acetate as the mobile phase and 20 mg injection with an
octadecyl column. The higher DP value oligomers were more difficult to
detect due to the fact that the weight percentage decreases with
increasing DP value.

4.4 Supercritical Fluid Chromatography (SFC)

The new technique of supercritical fluid chromatography (SFC) uses a


highly compressed gas as the mobile phase, with carbon dioxide as the
most widely used mobile phase in SFC. Since carbon dioxide is non-
polar, polar solutes require derivatization to enhance miscibility. Lafosse
et al. [34] in 1992 used an evaporative light scattering detector (ELSD)
344 MATTHEW J. MOLLAN JR. AND METIN CELIK

with both HPLC (to separate maltodextrin components) and SFC (to
separate sugars). Maltodextrin HPLC analysis was performed by
separation on octadecyl-bonded silica with a water-methanol gradient
and detection by ELSD.

4.5 N M R Spectrametry

The use of water-elimination Fourier transform nuclear magnetic


resonance to determine the degree of derivatization of maltodextrin with
acrylic acid glycidyl ester at alkaline pH was described by Lepisto, et al.
[3 51. They used a JEOL JNM-FX 100 FT-NMR spectrometer at 100
MHz and were able to determine the number of double bonds in
acryloylated maltodextrins.

Radosta ana Schierbaum [36]used NMR relaxation times (both TI and


T2) for starch polysaccharides in solution to characterize bound and free
water. German et al. 1371used Proton Pulse NMR to study water
mobility in solutions of maltodextrins and on the sol-gel transition mode.
They found that maltodextrin gels have a micro-inhomogeneous
structure, and were able to describe the maltodextrin gelation process by
percolation theory.

McIntyre and Vogel [38] used two dimensional NMR to obtain the
complete assignment of the overlapping proton NMR spectrum of
starches, as well as maltodextrins, in D 2 0 solutions at 25'C. Mora-
Gutierrez and Baianu [39] used solid-state I3CNMR techniques, and
found significant differences between the spectra of corn and potato
maltodextrins.

5. Acknowledgments

We would like to acknowledge A. Newman, G. Young, T. Cortina, and


H. Brittain of Bristol-Myers Squibb for their work for the TG, x-ray
powder diffraction, and surface area data. We would also like to thank
the Edward Mendell Company for their generous support of (MM)
throughout the research project.
MALTODEXTRIN 345

6. References

1) United States Pharmacopeia and National Formulary, NF XVII, Sixth


Supplement, Rockville, MD, pp. 2962-2963 (1 994).

2) Kanig, J., "Direct Compression Tabletting Composition and


Pharmaceutical Tablet Produced There From", U.S. patent
3873694, March 25,1975.

3) Kenyon, M.M. and Anderson, R.J., "Maltodextrins and Low-


Dextrose-Equivalence Corn Syrup Solids", Chapter two in ACS
Symp. Ser., m, pp. 7-1 1 (1988).

4) Maltrin@Maltodextrin, Product Data Technical Bulletin, Grain


Processing Corporation, Muscatine, Iowa.

5 ) Wartman, A.; Hagberg, C.; and Eliason, M., "Refractive Index-Dry


Substance Relationships for Commercial Corn Syrups", J. Chem.
Eng. Data, 2, 459-468 (1976).

6) Porter, S.C., and Woulicki, E.J., "Maltodextrin Coating", South


African patent ZA8S00209A, August 28,1985.

7) Porter, S.C., and Waznicki, E.J., "Maltodextrin Coating", South


African patent ZA5000209A, August 28,1985.

8) Lloyd, N.E., and Nelson, W.J., Starch: Chemistry and Tec-


Second Edition, R.J. Whistler, eds., Academic Press, London,
p. 611 (1984).

9) Schmidt, P., and Brogman, B., "Effervescent Tablets: Choice of a


New Binder for Ascorbic Acid", Acta Pharm. Technol., 34. 22-
26, (1988).

10) Visavarungroj, N., and Remon, J.P., "Evaluation of Maltodextrin as


a Binding Agent", Drug Dev. Ind. Pharm., 18,1691-1700
(1992).
346 MATTHEW J. MOLLAN JR.AND METIN CELIK

11 ) Parrott, E.L., "Comparative Evaluation of a New Direct


Compression Excipient, Soludex 15", Drug Dev. Ind. Pharm.,
B, 561-583 (1989).

12) Papadimitriou, E.; Efentakis, M.; and Choulis, N.H., "Evaluation of


Maltodextrins as Excipients for Direct Compression Tablets and
their Influence on the Rate of Dissolution", Int. J Pharm., 86,
131-136 (1992).

13) Mollan, M.J., and Celik, M., "Characterization of Directly


Compressible Maltodextrins Manufactured by Three Different
Processes", Drug Dev.Ind. Pharm., le , 2335-2358 (1993).

14) Munox-Ruiz,A.; Mondero Perales, M.C.; Velasco Antequera, M.V.,


and Jimenez-Castellanos, M.R., "Physical and Rheological
Properties of Raw Materials", S. T.P. Pharma Sci., I, 307-3 12
(1993).

15) Myers$., and Shively, M.L., "Solid-state Emulsions: The Effects of


Maltodextrin on Microcrystalline Aging", Pharm. Res., 14,
1389-1391 (1993).

16) Mollan, M.J.,


. . of a Roller C o w t e d Maltodextrin
for Direct Corny- * , PhD thesis, Rutgers, The State
University of New Jersey, May 1993.

17) Brunauer, S.; Emmett, P.; and Teller, E., "Adsorption of Gases,
Multimolecular Layers", J. Am. Chem. Soc., 612,309-316
(1938).

18) Ritter, H.C., and Drake, L.C., "Pore-Size Distribution in Porous


Materials. Pressure Porosimetry and Determination of Complete
Macropore Distribution", Ind. Eng. Chem. Anal. Ed. , 12, 782-
786 (1945).

19) Carr, R., "Evaluating Flow Properties of Solids", Chem. Eng.,


a(2), 163-168 (1965).
MALTODEXTRIN 347

20) Churms, S.C., "Recent Developments in the Chromatographic


Analysis of Carbohydrates", J. Chromatography, 5M, 555-583
(1990).

2 1) Hicks, K.B., and Hotchkiss, A.T., "High-Performance Liquid


Chromatography of Plant-Derived Oligosaccharides on a New
Cation-Exchange Resin Stationary Phase: HPX-22H",
JChromatography, 441, 382-386 (1988).

22) Lee, Y .C., "High-Performance Anion-Exchange Chromatography


for Carbohydrate Analysis", Anal. Biochem., .l.@
151-162,
(1990).

23) Covacevich, M.T., and Richards, G.N., Tontinuous Quantitative


Thin-Layer Chromatography of Oligosaccharides", J.
Chromatography, 129, 420-425 (1976).

24) Bosch-Reig, F.; Marcote, M.J.; Minana, M.D.; and Cabello, M.L.,
"Separation and Identification of Sugars and Maltodextrins by
Thin Layer Chromatography: Application to Biological Fluids
and Human Milk", Talanta, 3,1493-1498 (1992).

25) Vajda, J. and Pick, J., "Separation of some mono-, di-, tri-, and
oligosaccharides", 2ndProc. Int. Conf. Biochem. Sep., J. Pick
and J. Vajda, eds., 191-197 (1988).

26) Scobell, H.D.; Brobst, K.M.; and Steele, E.M., "Automated Liquid
Chomatographic System for Analysis of Carbohydrate Mixtures",
Cereal Chem., 54, 905-917 (1977).

27) Scobell, H.D., and Brobst, K.M., "Rapid High-Resolution


Separation of Oligosaccharides on Silver Form Cation Exchange
Resins", J. Chromatography, U ,5 1-64 (198 1).

28) Cheetham, N.W.H.; Sirimanne, P.; and Day, R.W., "High


Performance Liquid Chromatography Separation of Carbohydrate
Oligomers", J. Chromatography, 222,439-444 (198 1).
348 MATTHEW J. MOLLAN JR. AND METIN CELIK

29) Warthesen, J.J., "Analysis of Saccharides in Low Dextrose


Equivalent Starch Hydrolysates using High Performance Liquid
Chromatography", Cereal Chem., a, 194-195 (1984).

30) Brooks, J.R., and Griffin, V.K., "Saccharide Analysis of


Corn(maize) Syrup Solids and Maltodextrins Using High
Performance Liquid Chromatography", Cereal Chem., a,253-
255 (1987).

3 1) McGinnis, G.D.; Prince, S.; and Lowrimore, J., "The Use of


Reverse-Phase Columns for Separation of Unsubstituted
Carbohydrates", J. Carbohydrate Chem., 5, 83-97 (1 986).

32) Honda, S.; Akao, E.; Suzuki, S.; Okuda, M.; Kakehi, K.; and
Nakamura, J., "High Performance Liquid Chromatography of
Reducing Carbohydrates as Strongly Ultraviolet Absorbing and
Electrochemically Sensitive 3-methyl- 1-phenyl-5-pyrazolone
Derivatives'', Anal. Biochem., .€.@, 35 1-357 (1989).

3 3 ) Niessen, W .M.A.; Van der Hoeven, R.A.M.; Van der Greef, J.,
Schols, H.A., and Voragen, A.G.J., "Online Liquid
Chromatography/Thermospray Mass Spectrometry in the
Analysis of Oligosaccharides", Rapid Commun. Mass Spectrom.,
$, 197-202 (1992).

34) Lafosse, M.; Elfakir, C.; Morin-Allory, L.; and Dreux, M.,
"Advantages of Evaporative Light Scattering Detection in
Pharmaceutical Analysis by High Performance Liquid
Chromatography and Supercritical Fluid Chromatography", J
High Resolution Chromatogr., 15, 3 12-3 18 (1992).

35) Lepisto, M.; Artursson, P.; Edman, P.; Laakso, T.; and Sjoholm, I.,
"Determination of the Degree of Derivatization of Acryloylated
Polysaccharides by Fourier Transform Proton NMR
Spectroscopy", Anal. Biochem., U, 132-135 (1983).

36) Radosta, S. and Schierbaum, F., "Polymer-Water Interaction of


Maltodextrins Part II.", Starch, a, 428-430 (1989).
MALTODEXTRIN 349

37) German, M.L., Blumenfeld, A.C., Yuryev, V.P., and Tolstoguzov,


V.B., "AnNMR Study of Structure Formation in Maltodextrin
Systems", Carbohydrate Polymers, U, 139-146 (1989).

38) McIntyre, D.D. and Vogel, H.J., "Two-Dimensional Nuclear


Magnetic Resonance Studies of Starch and Starch Products",
Starch, 42, 287-293 (1990).

39) Mora-Gutierrez, A. and Baianu, I.C., "Carbon-13 Nuclear Magnetic


Resonance Studies of Chemically Modified Waxy Maize Starch,
Corn Syrups, and Maltodextrins. Comparisons with Potato
Starch and Potato Maltodextrins", J. Agric. Food Chem., 22,
1057-1062 (1991).

You might also like