The Effects Of Adulterants And Selected Ingested Compounds On Drugs-Of-Abuse Testing

Jacob Bell

New Member
The Effects of Adulterants and Selected Ingested
Compounds on Drugs-of-Abuse Testing in Urine

Amitava Dasgupta, PhD
Key Words: Adulterants; Nitrite; Potassium chlorochromate; Drugs of abuse; Poppy seed cake; Hemp oil

A b s t r a c t
Household chemicals such as bleach, table salt,
laundry detergent, toilet bowl cleaner, vinegar, lemon
juice, and eyedrops are used for adulterating urine
specimens. Most of these adulterants except eyedrops
can be detected by routine specimen integrity tests
(creatinine, pH, temperature, and specific gravity);
however, certain adulterants such as Klear, Whizzies,
Urine Luck, and Stealth cannot. These adulterants can
successfully mask drug testing if the concentrations of
certain abused drugs are moderate. Several spot tests
have been described to detect the presence of such
adulterants in urine. Urine dipsticks are commercially
available for detecting the presence of such adulterants,
along with performance of tests for creatinine, pH, and
specific gravity. Certain hair shampoo and salivacleaning
mouthwashes are available to escape
detection in hair or saliva samples, but the effectiveness
of such products in masking drugs-of-abuse testing has
not been demonstrated. Ingestion of poppy seed cake
may result in positive screening test results for opiates,
and hemp oil exposure can cause positive results for
marijuana. These would be identified as true-positive
results in drugs-of-abuse testing even though they do
not represent the actual drug of abuse.
Persons abusing drugs attempt to adulterate urine specimens
to escape detection in drug testing. In this review, the
effects of diluted urine, household chemicals, and Internetbased
urinary adulterants such as potassium nitrite, pyridinium
chlorochromate, and glutaraldehyde on urine drug testing
will be discussed. Moreover, techniques available in clinical
laboratories to identify such adulterated specimens, including
spot tests and dipsticks, are also addressed. The effect of
ingesting poppy seed cakes, hemp oil, and coca tea on analysis
for drugs of abuse is addressed.
Drug abuse is a critical problem not only in the United
States but also throughout the world. Commonly abused
drugs are cocaine, cannabinoids, amphetamine, phencyclidine
(PCP), and benzodiazepines. Designer drugs such as
3,4-methylenedioxyamphetamine and 3,4-methylenedioxymethamphetamine
(Ecstasy) are commonly used in rave
parties, along with Rohypnol (flunitrazepam) and γ-hydroxybutyric
acid.
On September 15, 1986, President Reagan issued
Executive Order No. 12564 directing federal agencies to
achieve a drug-free work environment. The Department of
Health and Human Services (formerly the responsibility of the
National Institute on Drug Abuse) developed guidelines and
protocols for drugs-of-abuse testing. The Mandatory
Guidelines for Federal Workplace Drug Testing Programs
were first published in the Federal Register on April 11,
1988,1 and were since revised on June 9, 1994,2 and
September 30, 1997.3 Another notice was issued on April 13,
2004.4 The overall testing process under mandatory testing
consists of proper specimen collection, initiation of the chain
of custody, and analysis of the specimen (screening and gas
chromatography—mass spectrometry [GC-MS] confirmation,
if needed) by a Substance Abuse and Mental Health Services
Administration (SAMHSA)-certified laboratory. The screening
by immunoassay should be performed using a US Food
and Drug Administration—approved method. The confirmation
should be performed by a second technique, preferably by
GC-MS). Federal guidelines for cutoff levels of 5 abused
drugs in screening and confirmation phases of the drug testing
program are summarized in ❚Table 1❚.
It is estimated that approximately 20 million employees
are screened each year in the United States for illicit drugs.
Marijuana is the most frequently abused drug. Drug testing
programs in the United States can be classified as mandatory
or nonmandatory. In the mandatory category (eg, the
Department of Transportation program), a regulated employer
is required by federal regulation to test the employees. In the
nonmandatory category, employers choose to test for reasons
other than federal requirements. Private employers who are
not mandated to test under federal authority have instituted
employee drug testing to create a drug-free workplace. These
programs also formalized the role of a specialist physician
termed medical review officer (MRO). The MRO is an integral
part of a drug testing program who can determine the cause of
positive results in drug testing (eg, interference or other prescription
drugs) and counsel the employee. It is required that
a laboratory submit a drug testing result to the MRO within 5
working days of receiving the specimen, and the result must
be certified by the certifying scientist.
Federal guidelines define an adulterated specimen as a
urine specimen containing a substance that is not a normal constituent
or containing an endogenous substance at a concentration
that is not a normal physiologic concentration. In the military
where the urine sample collection process is supervised,
the chances of receiving adulterated specimens are reduced, but
in preemployment screening where direct supervision of specimen
collection is not practiced, a person may attempt to escape
detection by adulterating specimens to avoid unwanted consequences
of failing a drug test. Several precautions are taken by
personnel at the collection site to avoid adulteration of submitted
specimens, such as asking the donor to remove outer garments
(coat or jacket) that may contain concealed adulterating
substances. The collector needs to ensure that all personal
belongings such as purse or briefcase stay with the collector,
but the donor may retain his or her wallet. The collector also
directs the donor to empty his or her pockets and display the
items to ensure that no item is present that could be used to
adulterate the specimen.
Commercially Available Products to
Escape Detection
Common household chemicals such as laundry bleach,
table salt, toilet bowl cleaner, hand soap, and vinegar have
been used for many years as adulterants of urine specimens in
an attempt to avoid a positive drug test result. There is also a
popular belief that drinking goldenseal tea helps escape detection.
Varieties of products have become commercially available
and can be ordered through Internet sites and toll-free
numbers. Home test kits are also available commercially to
test for certain drugs. Synthetic urine is available from Internet
sites as a sure method to beat a drug test in settings where collection
of a urine specimen is not supervised. Quick Fix
Synthetic Urine is a bottle of premixed urine with all the characteristics
of natural urine (correct pH, specific gravity, and
creatinine level). The product can be heated in a microwave
oven for up to 10 seconds to achieve a temperature between
90°F and 100°F. It can also be taped next to a heating pad to
maintain the normal temperature of urine for up to 6 hours in
a pocket.
Commercially available products to beat drug tests can be
classified under 2 broad categories. The first category includes
specific fluids or tablets, along with substantial water intake,
to flush out drugs and metabolites. Many of these products can
produce diluted urine, and the concentrations of drugs or
metabolites can be significantly reduced. Common products
are Absolute Detox XXL Drink, Absolute Carbo Drinks,
Ready Clean Drug Detox Drink, Fast Flush Capsules, and
Ready Clean Gel Capsules. The second category of products
includes in vitro urinary adulterants, which are added to urine
after specimen collection to pass a drug test. Stealth (contains
peroxidase and peroxide), Klear (nitrite), Instant Clean ADDIT-
ive (glutaraldehyde), and Urine Luck (pyridinium
chlorochromate [PCC]) are urinary adulterants available
through the Internet. Iodine is a strong oxidizing agent and
may potentially destroy abused drugs, especially tetrahydrocannabinol
(THC), if present in urine.5 A recent article indicates
that papain, a cysteine protease with intrinsic ester
hydrolysis ability, can significantly reduce the concentration
of 11-norcarboxy-Δ9-tetrahydrocannibinol (THC-COOH), a
metabolite of marijuana, if added to the urine specimen in
vitro. This product is relatively inexpensive and commercially
available. Papain, however, did not significantly decrease concentrations
of other drugs analyzed (by enzyme multiplied
immunoassay and fluorescence polarization immunoassay
[FPIA]), except nordiazepam.6
Diluted Urine: An Attempt to Escape
Detection by Decreasing the
Concentrations of Drugs of Abuse
A negative result for the presence of abused drugs in a
urine specimen does not mean that no drug was present. It is
possible that the amount of drug was below the cutoff value
for detection in the laboratory assay. Diluting urine is a simple
way to make an otherwise positive drug test result negative.
Federal guidelines recommend placing a toilet bluing agent in
the toilet tank, if possible, so that the reservoir of water in the
toilet bowl always remains blue. There should be no other
source of water in the enclosure where urination takes place.
Consumption of a large amount of fluid before drug testing
is a way to avoid a positive test result.7 A creatinine concentration
below 20 mg/dL or a specific gravity below 1.003
should be considered an indication of diluted urine. Creatinine
analysis in urine is a very effective method to detect diluted
urine. Needleman and Porvaznik8 considered a creatinine
value of less than 10 mg/dL as suggestive of replacement of a
urine specimen largely by water. Beck et al9 reported that 11%
of all urine specimens submitted to their laboratory for drugsof-
abuse testing were diluted.
The SAMHSA program does not allow analysis of diluted
urine specimens at lower screening and confirmation cutoffs.
However, the Correctional Services of Canada (CSC), for
diluted urine specimens, incorporates lower screening and
confirmation cutoffs for drugs and metabolites (amphetamine:
screening cutoff, 100 ng/mL; confirmation cutoff, 100 ng/mL;
benzoylecgonine: screening and confirmation cutoffs, 15
ng/mL; opiates, screening and confirmation cutoffs, 120
ng/mL; PCP: screening and confirmation cutoffs, 5 ng/mL;
and cannabinoids, screening cutoff, 20 ng/mL; confirmation
cutoff, 3 ng/mL). Fraser and Zamecnik10 reported that 7,912
urine specimens collected and analyzed between 2000 and
2002 by the CSC were diluted, and of those specimens, 26%
had positive screening results using SAMHSA cutoff values.
When lower values for cutoff and confirmation were adopted,
1,100 specimens tested positive for 1 or more illicit drugs. The
positive rate of diluted specimens was 18.2% in CSC institutes
and 22.3% in parolee specimens. The drug most often confirmed
as positive in diluted specimens was marijuana.
Codeine and/or morphine were also commonly confirmed in
these urine specimens and ranked second after marijuana in
prevalence.10 Soldin et al11 reported that there was more than
a 100% increase in cocaine-positive specimens when the cutoff
was lowered to 80 ng/mL from 300 ng/mL in a pediatric
population because neonates cannot concentrate urine to the
same extent as adults.
Luzzi et al12 studied the analytic performance criteria of
3 immunoassay systems (EMIT, Beckman EIA, and Abbott
FPIA [Abbott Laboratories, Abbott Park, IL]) for detecting
abused drugs with concentrations less than established cutoff
values. They concluded that drugs can be screened at concentrations
much lower than the cutoffs established by SAMHSA.
For example, the authors proposed a THC-COOH cutoff value
of 35 ng/mL using EMIT and 14 ng/mL for the Beckman EIA
and the FPIA, for which SAMHSA guidelines stated a cutoff
value of 50 ng/mL. The proposed cutoff values were based on
the studies of precision of the assays at proposed lower detection
limits at which the coefficient of variation was less than
20%. This lowering of the cutoff values increased the number
of positive specimens in the screening tests to 15.6%. A 7.8%
increase was also observed in the confirmation stage of drugsof-
abuse testing.12
New SAMHSA regulations indicate that a specific gravity
less than 1.0010 (new refractometers can detect such low
concentrations, to 4 decimal places accurately) or more than
1.020 and a creatinine concentration less than 5 mg/dL are
inconsistent with normal human urine.13 Edgell et al14 performed
a controlled hydration study with 56 volunteers to
determine whether it is possible to produce such diluted urine.
Subjects were given 2,370 mL of fluid, and urine specimens
were collected at the end of each hour for a 6-hour test period.
No urine specimen satisfied the paired substitution criteria
(specific gravity ≤1.001 or >1.020 and creatinine ≤5.0 mg/dL)
for diluted urine (although 55% of subjects produced at least
1 diluted urine specimen during the first 3 hours of hydration
with a creatinine concentration <20 mg/dL and a specific
gravity <1.003). This finding supports the criteria set by
SAMHSA for classifying a specimen as substituted.14
Barbanel et al15 studied specific gravity and/or creatinine
concentrations in 803,130 random urine specimens submitted
to the laboratory. Creatinine and specific gravity measurements
were performed on 13,467 of these specimens, and
none of them met the lower limit for specific gravity (<1.001)
and creatinine concentration (<5 mg/dL). The samples that
met 1 of the 2 criteria were from neonates or patients who
were severely ill and unlike anyone in the workforce undergoing
testing for abused drugs. Samples from 11 patients met the
criteria for substituted urine (creatinine concentration, <5
mg/dL; specific gravity, >1.020), but all of the patients were
seriously or terminally ill.15 Cook et al16 demonstrated that an
osmolality cutoff of less than 50 mOsm/kg (50 mmol/kg) can
be indicative of substituted urine.
Flushing, Detoxification Agents,
Diuretics, and Herbal Tea to Escape
Detection
Flushing and detoxification are frequently advertised as
effective means of passing drug tests. Cone et al17 evaluated
the effect of excess fluid ingestion on false-negative marijuana
and cocaine urine test results. They studied the ability of
Naturally Clean Herbal Tea, goldenseal tea, and
hydrochlorothiazide to cause false-negative results. Volunteers
drank 1 gallon of water (4 doses in a 4-hour period) or herbal
tea or hydrochlorothiazide 22 hours after smoking marijuana
cigarettes or intranasal administration of cocaine. The creatinine
levels dropped below the cutoff 2 hours after intake of
excessive fluid. Marijuana and cocaine metabolite levels (as
measured by EMIT and TDx) decreased significantly, and
results frequently switched from positive to negative after subjects
consumed 2 quarts of fluid. Even excess water consumption
effectively diluted a urine specimen to cause false-negative
results. Consumption of herbal tea produced diluted urine
faster than consumption of water alone.17 Consumption of
goldenseal tea produces dark urine and can be identified by
visual inspection.18 A more sophisticated approach to identify
marker compounds of goldenseal tea in a specimen suspected
of contamination is the use of high-performance liquid chromatography.
19
Diuretics are used in sports for 2 purposes: to flush out
previously taken banned substances by forced diuresis and to
achieve quick weight loss to qualify for a group in a lower
weight class. Ingestion of salicylate-containing drugs and
sodium bicarbonate is also done to avoid positive results in
drug testing.6 The Medical Commission of the International
Olympic Committee bans the use of diuretics. There is no
commercially available immunoassay for detecting diuretics
such as hydrochlorothiazide in urine, and a sophisticated technique
such as liquid chromatography combined with tandem
mass spectrometry is necessary to confirm the presence of
diuretics in doping analysis.20
Common Household Chemicals as
Urinary Adulterants
People try to beat drug testing by adding adulterants to
urine specimens. Several adulterants can cause false-negative
results in drug testing by immunoassays. Common adulterants
for masking drug testing are as follows: (1) table salt, (2)
household vinegar, (3) liquid laundry bleach, (4) concentrated
lemon juice, (5) goldenseal tea (produces dark urine), and (6)
eyedrops.
Although FPIA is less subject to interference from adulterants
than the EMIT assay, some interference has also been
reported with FPIA. Sodium chloride caused negative interference
with all drugs tested by EMIT and a slight decrease in
measured concentrations of benzodiazepines by FPIA.
Sodium bicarbonate caused false-positive results with an
EMIT opiate assay and with PCP testing by FPIA. Hydrogen
peroxide caused false-positive benzodiazepine results by
FPIA.21 Uebel and Wium22 studied the effects of household
chemicals (sodium hypochlorite, Dettol [chloroxylenol], glutaraldehyde,
Pearl hand soap, ethanol, isopropanol, and peroxide)
on cannabis and methaqualone test results using EMIT
assays. Most of the agents tested interfered with the test
results, and the greatest effect was observed with glutaraldehyde
and Pearl hand soap for methaqualone (false-negative).
Dettol and Pearl hand soap also caused false-negative results
in cannabis tests. Addition of isopropanol, ethanol, and peroxide
invalidated a methaqualone test result.22
Schwarzhoff and Cody23 studied the effect of 16 adulterating
agents (ammonia-based cleaner, L-ascorbic acid, eyedrops,
drain opener, goldenseal tea, lemon juice, lime solvent,
liquid bleach, liquid hand soap, methanol, sodium chloride,
tribasic potassium phosphate, toilet bowl cleaner, white vinegar,
ionic detergent, and whole blood anticoagulated with
EDTA) on analysis of urine by FPIA for abused drugs. They
tested these adulterating agents at a 10% by volume concentration
of urine with the exception of goldenseal tea because
of its insolubility. For goldenseal tea, 1 capsule was suspended
in 60 mL of urine. Of 6 drugs tested (cocaine metabolites,
amphetamines, opiates, PCP, cannabinoid, and barbiturates),
the cannabinoid test was most susceptible to adulteration.
Approximately half of the agents tested (ascorbic acid, vinegar,
bleach, lime solvent, eyedrops, and goldenseal tea) caused
false-negative results. Cannabinoid and opiate assays were
susceptible to bleach, and actual degradation of THC was confirmed
by GC-MS. The PCP and benzoylecgonine (the
metabolite of cocaine) analyses were affected by alkaline
agents.23 Baiker et al24 reported that hypochlorite (a common
ingredient of household bleach) adulteration of urine caused a
decreased concentration of THC as measured by GC-MS. A
false-negative result was also observed with the FPIA screen
and the Roche Abuscreen (Roche Diagnostics, Indianapolis,
IN).24 Another report described adulteration of urine specimens
with denture cleaning tablets.25
The ability of eyedrops to cause false-negative drug test
results in the screening phase of the analysis is troublesome
because the presence of components of eyedrops in adulterated
urine cannot be detected by routine specimen integrity testing
or routine urine analysis. Pearson et al26 studied in detail
the effect of eyedrops on drugs-of-abuse testing and the mechanism
by which components of eyedrops produce false-negative
drug testing results. Eyedrops are effective in causing
false negative results in the analysis of the THC metabolite,
THC-COOH. GC-MS analysis showed that there was no
modification in the structure of the THC metabolite by the
components of eyedrops. At low concentrations of eyedrops,
the false-negative cannabinoid result was due to the benzalkonium
chloride ingredient of eyedrops. Eyedrops decreased the
THC assay results in EMIT/drugs-of-abuse assays and
Abuscreen, although eyedrops had no effect on the glucose-6-
phosphate dehydrogenase drug conjugate used in the EMIT
assay. Results of the ultrafiltration studies with eyedrops suggest
that the THC metabolite partitions between the aqueous
solvent and the hydrophobic interior of benzalkonium chloride
micelles, thus reducing the availability of THC metabolite
in antibody-based assays.26 Eyedrops and analgesic heat rub
ointment can also cause false-negative drug test results with
sweat testing.27 Components of eyedrops in urine may be
detected by using high-performance liquid chromatography
combined with UV detection at 262 nm, a method originally
developed for analysis of ophthalmic formulations.28
Specimen Integrity Tests
The collection site and the laboratory have a number of
mechanisms to detect potentially invalid specimens. The temperature
of the urine, for example, should be 90.5°F to 98.9°F.
The specific gravity should be between 1.005 and 1.030, and
the pH should be between 4.0 and 10.0. The creatinine concentration
should be 20 to 400 mg/dL. A specimen is considered
diluted if the creatinine is less than 20 mg/dL and the specific
gravity is less than 1.003. The laboratory should perform
creatinine and pH analysis of all specimens submitted for
drugs-of-abuse testing. Additional tests are also recommended
to detect the presence of other adulterants.
Determination of specific gravity is mandatory for any
specimen with a creatinine concentration of less than 20
mg/dL. Substituted urine specimens have creatinine concentrations
less than 5 mg/dL and a specific gravity of less than
1.001 or more than 1.020. The urine is adulterated if the pH is
less than 3 or more than 11.6 Adulteration with sodium chloride
at a concentration necessary to produce a false-negative
result always produces a specific gravity of more than 1.035.
Use of household chemicals such as bleach, acid, soap, detergent,
and vinegar alters the pH of urine to a value outside the
physiologic range that can be easily detected by specimen
integrity tests. Specimens adulterated with liquid soap are usually
cloudy. The presence of eyedrops in adulterated urine
cannot be detected by routine specimen integrity testing.
Moreover, newer urine adulterants such as Urine Luck,
UrinAid, Klear, and Whizzies can also cause false-negative
result in drug tests. The presence of these compounds in urine
may escape detection by routine specimen integrity tests.
Adulteration Product Urine Luck
Wu et al29 reported that the active ingredient of Urine
Luck is 200 mmol/L of PCC. They reported a decrease in the
response rate for all EMIT II drug screens and for the
Abuscreen morphine and THC assays. In contrast, Abuscreen
amphetamine assays produced a higher response, and no
effect was observed on the results of benzoylecgonine and
PCP. This adulteration of urine did not alter GC-MS confirmation
test results for methamphetamine, benzoylecgonine, and
PCP. However, apparent concentrations of opiates and THC as
determined by GC-MS were reduced.29 Paul et al30 also studied
the effect of Urine Luck on testing for drugs of abuse.
When THC-COOH—containing urine specimens were treated
with 2 mmol/L of PCC, 58% to 100% of the THC-COOH was
lost. The loss increased with decreasing pH and increasing
time of incubation (0-3 days). There was no effect on the concentration
of free codeine or free morphine if the pH of the
urine was in the range of 5 to 7, but at a lower pH, significant
loss of free morphine was observed. Amphetamine, methamphetamine,
benzoylecgonine, and PCP remained unaffected
by PCC at a urine pH of 3 to 7.30
Spot Tests for Detecting Urine Luck (PCC) in Urine
Wu et al29 also described the protocol for detection of
PCC in urine using spot tests. The indicator solution contains
10 g/L of 1,5-diphenylcarbazide (DPC) in methanol. The indicator
detects the presence of the chromium ion and is colorless
when prepared. Two drops of indicator solution is added
to 1.0 mL of urine. If the sample becomes reddish purple, the
test result is positive.29
Paul et al30 also used DPC for detection of PCC in urine.
When this reagent was added, the sample immediately
became red-violet if PCC was present. The chromium-DPC
complex showed a characteristic absorption peak at 544 nm
and a shoulder peak at 575 nm. The ratio of absorption can be
used to detect the presence of PCC as chromium in urine, and
the concentration of chromium can be estimated by measuring
absorption at 544 nm, with a linear association between concentrations
of 0.5 and 20 μg/mL.30 Addition of a few drops of
PCC-adulterated urine to approximately 0.5 mL of potassium
iodide solution followed by addition of a few drops of 2N
hydrochloric acid leads to liberation of iodine from potassium
iodide, and this can also be used as a spot test to detect PCC.
Addition of 4 or 5 drops of 3% hydrogen peroxide to approximately
200 μL of urine adulterated with PCC (approximately
6-7 drops from a transfer pipette) caused rapid formation of a
dark brown color (due to reduction of heptavalent chromium
by hydrogen peroxide), and a dark brown precipitate appeared
on standing. In contrast, unadulterated urine turned colorless
after addition of hydrogen peroxide.31
Ferslew et al32 described a capillary ion electrophoresis
technique for detecting chromate ion and nitrite ion in urine
specimens suspected of adulteration. The DPC colorimetric
test for chromate, which can be easily automated, can serve as
a screening test. Capillary electrophoretic analysis can be used
to confirm the presence of chromate in adulterated specimens,
if necessary. A good correlation was observed between chromate
concentrations in urine using the colorimetric test and
the capillary electrophoretic analysis.33 Paul34 described 6
spectroscopic methods for the detection of oxidants, including
chromate. The presence of oxidants (as adulterants in urine)
was established by initial oxidation of ferrous to ferric ion and
then detection of ferric ion by chromogenic oxidation or complex
formation. Paul34 used N,N-dimethylamino-1,4-
phenylenediamine, 2,2'-azino-bis(3-ethylbenzthiazoline-6-
sulfonic acid), and 2-amino-p-cresol for chromogenic oxidation.
The reagents for the chromogenic complex formation
were xylenol orange, 8-hydroxy-7-iodo-5-quinolinesulfonic
acid, and 4,5-dihydroxy-1,3-benzene-disulfonic acid.34
Adulteration of Urine With Nitrite-
Containing Agents
The product Klear comes in 2 microtubes containing 500
mg of white crystalline material. This product readily dissolves
in urine without affecting color or temperature. Klear
may cause a false-negative GC-MS confirmation result for
marijuana. ElSohly et al35 first reported this product as potassium
nitrite and provided evidence that nitrite leads to decomposition
of ions of 9-THC and its internal standard. They
reported that using a bisulfite step at the beginning of sample
preparation could eliminate such interference.35
Tsai et al36 further studied the effects of nitrite on
immunoassay screening for other drugs, including cocaine
metabolites, morphine, THC metabolites (THC-COOH),
amphetamine, and PCP. Nitrite at a concentration of 1.0 mol/L
had no effect on the Abuscreen assay. At a higher nitrite concentration,
the amphetamine assay became more sensitive,
and the THC metabolite assay became less sensitive. The GCMS
analyses of benzoylecgonine, morphine, amphetamine,
and PCP were not affected, whereas recovery of the THC
metabolite was significantly reduced. Again, this interference
could be eliminated by bisulfite treatment.36
Duration of nitrite exposure and the urine matrix also
affect the THC-COOH assay. In an in vitro study, 40 clinical
urine specimens confirmed as positive for THC-COOH were
supplemented with 1.15 or 0.30 mol/L of nitrite. The results
indicated that the pH of the urine and the original drug concentrations
have major roles in dictating the effectiveness of
nitrite in causing false-negative THC metabolite test results.
With an acidic pH, significant decreases in the immunoassay
screening results can be observed in all urine specimens within
4 hours of adulteration with nitrite regardless of the original
concentration of THC-COOH (range of concentration,
33-488 ng/mL as determined by GC-MS). All specimens
were negative for THC-COOH after 1 day. In contrast, the
immunoassay results of urine specimens with a basic or neutral
pH were less affected by nitrite exposure. Approximately
two thirds of the samples with pH values greater than 7.0 had
positive immunoassay results, even 3 days after supplementing
with nitrite.37
Nitrite in urine may arise in vivo and is found in urine in
a low concentration. Patients receiving medications such as
nitroglycerine, isosorbide dinitrate, nitroprusside, and ranitidine
may have increased nitrite levels in their blood. However,
concentrations of nitrite were less than 36 μg/mL in specimens
cultured positive for microorganisms, and nitrite concentrations
were less than 6 μg/mL in patients receiving medications
that are metabolized to nitrite. On the other hand, nitrite concentrations
were 1,910 to 12,200 μg/mL in urine specimens
adulterated with nitrite.38 Whizzies is another urine adulterant
available from the Internet. This adulterant also contains
potassium nitrite.
Spot Tests for Nitrite
Addition of a few drops of a nitrite-adulterated urine
specimen to 0.5 mL of 1% potassium permanganate solution
followed by addition of a few drops of 2N hydrochloric acid
turns a pink permanganate solution colorless with effervescence.
The presence of very high glucose in urine (>1,000
mg/dL) and ketone bodies may cause a false-positive result.
However, it takes approximately 2 to 3 minutes for the solution
to turn colorless. On the other hand, if nitrite is present,
the solution turns colorless immediately. Another spot test to
detect nitrite uses a 1% potassium iodide solution. Addition of
a few drops of nitrite-adulterated urine to 0.5 mL of potassium
iodide solution followed by addition of a few drops of 2N
hydrochloric acid results in immediate release of iodine from
the colorless potassium iodide solution. Shaking of this solution
with n-butanol results in the transfer of iodine to the
organic phase. If no nitrite is present, the potassium iodide
solution remains colorless. There is no interference from a
high glucose level or ketone bodies if present in the urine.31
Nitrite can also be detected by diazotizing sulfanilamide
and coupling the product with N-(1-napthyl)ethylenediamine.
The presence of nitrite in urine can also be confirmed by
analysis using high-performance liquid chromatography using
an IonPac AS 14 analytic column with an anion self-generating
suppressor and conductivity detector. By using a single
point calibration, the assay was linear up to a nitrite concentration
of 12,000 μg/mL. The detection limit was 30 μg/mL.39
Kinkennon et al40 described a capillary electrophoresis
method for detection of nitrite in urine specimens suspected of
adulteration. The method involves separation of nitrite by capillary
electrophoresis and direct UV detection at 214 nm.
Separation was achieved by using a bare fused silica capillary
column and a 25-mmol/L phosphate buffer at pH 7.5. The
method was linear for a nitrite concentration of 80 to 1,500
μg/mL, with a limit of detection of 20 μg/mL. However,
CrO4
2— and S2O8
2—, as well as high concentrations of Cl—,
interfered with the chromatography results.40
Stealth as a Urinary Adulterant
Stealth is an adulterant advertised as an effective way to
escape detection in a urine drug test. Stealth consists of 2
vials, one containing a powder (peroxidase) and another
containing a liquid (hydrogen peroxide). Both products are
added to the urine specimen. Stealth is capable of producing
false-negative results using the Roche ONLINE and CEDIA
(Microgenics, Fremont, CA) immunoassay methods when
marijuana metabolites, lysergic acid diethylamide, and opiates
(morphine) are present in the urine at 125% to 150% of
cutoff values. Adulteration of an authentic positive sample
provided by a marijuana user caused the sample to screen as
negative with these immunoassay reagents.41 A low concentration
of morphine (2,500 ng/mL) could be effectively
masked by Stealth, but a higher concentration (6,000
ng/mL) tested positive by immunoassay (Roche ONLINE
and Microgenics CEDIA). GC-MS confirmation can be
affected if Stealth is present in the urine. Cody et al42 reported
that GC-MS analysis of Stealth-adulterated urine using
standard procedures proved unsuccessful in several cases,
and in 4 of 12 cases, neither the drug nor the internal standard
was recovered.
Valtier and Cody43 described a rapid color test to detect
the presence of Stealth in urine. The addition of 10 μL of
urine to 50 μL of tetramethylbenzidine working solution followed
by addition of 500 μL of a 0.1-mol/L phosphate buffer
solution caused a dramatic color change of the specimen to
dark brown. Peroxidase activity could also be monitored by
using a spectrophotometer. A routine specimen integrity
check using pH, creatinine, specific gravity, and temperature
did not detect the presence of Stealth in urine.43 My
experience shows that if a few drops of a urine specimen
adulterated with Stealth are added to potassium dichromate
followed by a few drops of 2N hydrochloric acid, the specimen
becomes deep blue immediately, but the color usually
fades with time.
Glutaraldehyde as an Adulterant of Urine
Glutaraldehyde has also been used as an adulterant to
alter urine drug test results.44 This product is available
under the trade name UrinAid. The manufacturer sells this
product for $20 to $30 per kit. Each kit contains 4 to 5 mL
of glutaraldehyde solution, which is added to 50 to 60 mL
of urine. Glutaraldehyde solutions are available in hospitals
and clinics as a cleaning or sterilizing agent. A 10% solution
of glutaraldehyde is available from pharmacies as
over-the-counter medication for treatment of warts.
Glutaraldehyde at a concentration of 0.75% by volume can
lead to false-negative screening results for a cannabinoid test
using the EMIT II drugs-of-abuse screen. Amphetamine,
methadone, benzodiazepine, opiate, and cocaine metabolite
tests can be affected at glutaraldehyde concentrations
between 1% and 2% with EMIT immunoassays. At a concentration
of 2% by volume, the assay of cocaine metabolites
is significantly affected (apparent loss of 90% sensitivity).
A loss of 80% sensitivity was also observed with the
benzodiazepine assay.
Wu et al45 reported that glutaraldehyde also interfered
with the CEDIA for screening of abused drugs. Goldberger
and Caplan46 reported that glutaraldehyde caused false-negative
results with EMIT but also caused false-positive PCP
results with the FPIA (Abbott Laboratories) and Kinetic
Interaction of Microparticles in a Solution Immunoassay
(KIMS, Roche Diagnostics).
Although the presence of glutaraldehyde as an adulterant
in urine can be detected by GC-MS, Wu et al47 described a
simple fluorometric method for the detection of glutaraldehyde
in urine. When 0.5 mL of urine was heated with 1.0 mL
of a 7.7-mmol/L concentration of potassium dihydrogen phosphate
(pH 3.0) saturated with diethylthiobarbituric acid for 1
hour at 96°C to 98°C in a heating block, a yellow-green fluorophore
developed if glutaraldehyde was present. Shaking the
specimen with n-butanol resulted in the transfer of this adduct
to the organic layer, which can be viewed under long wavelength
UV light. Glutaraldehyde in urine can also be estimated
by using a fluorimeter.47
Mechanism of Action of Adulterants
Adulterants such as bleach cause false-positive results in
THC radioimmunoassays but false-negative results with FPIA
and the EMIT assay. These erroneous results are due to the
direct effect of bleach on the reagents in the immunoassays.
18,48 Adulterants that are strong oxidizing agents such as
Klear (potassium nitrite), Urine Luck (PCC), and Stealth (peroxidase
and hydrogen peroxide) cause false-negative results in
the immunoassays used for screening drugs by directly
destroying THC metabolites (THC-COOH). In the GC-MS
confirmation stage, these adulterants interfere with confirmation
of THC-COOH because of the destruction of THC-COOH
and the internal standard, as well as interference during the
extraction phase. To overcome this problem, the use of reducing
agents such as sodium hydrosulfite or sulfamic acid before
extraction has been recommended.35 However, such steps can
allow detection of the remaining THC-COOH but cannot
recover the lost concentration of the marijuana metabolite.
Most oxidizing agents used as adulterants are more effective
if the pH of urine is acidic. To prevent destruction of drugs
by oxidizing agents, addition of carbonate as a buffering agent
before or after the urine is voided has been recommended.49
Other oxidizing agents such as potassium permanganate,
hydrogen peroxide—ferrous ammonium sulfate, periodic acid,
potassium persulfate, and sodium oxychloride can also
destroy THC-COOH within 48 hours. The effect of oxidizing
agents on THC-COOH primarily depends on the reduction
potential, pH, temperature, time of reaction, and urine constituents.
Horseradish-peroxidase with hydrogen peroxide and
a combination of hydrogen peroxide with Japanese radish,
black mustard seed, and red radish are effective in destroying
THC-COOH. Interestingly, hydrogen peroxide alone was not
effective in destroying any drug.50
Adulterants can also interfere with the extraction process.
Stealth is known to interfere with extraction of codeine and
morphine for GC-MS confirmation.41 My experience indicates
that PCC is effective for decreasing the semiquantitative
response rate for THC and opiates using Abuscreen (FPIA).
The incubation time had an important role in decreasing the
response rate. Nitrite is very effective in reducing the response
rate of THC, but the PCP assay was also affected.
Federal Guidelines for Additional Testing
to Detect Adulterants
SAMHSA guidelines require additional tests for urine
specimens with abnormal physical characteristics or that show
characteristics of an adulterated specimen during initial
screening or confirmatory tests (eg, nonrecovery of internal
standard and unusual response). A pH less than 3 or more than
11 and a nitrite concentration greater than 500 μg/mL indicate
the presence of adulterants. A nitrite colorimetric test or a general
oxidant colorimetric test can be performed to identify
nitrite. These criteria are summarized in ❚Table 2❚.
Similarly, the presence of chromium can be confirmed by
a chromium colorimetric test or a general test for the presence
of oxidant. A confirmatory test can be performed using multiwavelength
spectrophotometry, ion chromatography, atomic
absorption spectrophotometry, capillary electrophoresis, or
inductively coupled plasma mass spectrometry.
The presence of halogen (chloride, fluoride, or bleach)
should be confirmed by a halogen colorimetric test or a general
test for the presence of oxidants. Confirmatory tests may
use multiwavelength spectrophotometry, ion chromatography,
atomic absorption spectrophotometry, capillary electrophoresis,
or inductively coupled plasma mass spectrometry.
The presence of glutaraldehyde should be detected by a
general aldehyde test or the characteristic immunoassay
response in 1 or more drug immunoassay tests for initial
screening. The presence of PCC should be confirmed by using
a general test for the presence of oxidant and a GC-MS confirmatory
test.
The presence of a surfactant should be verified by using
a surfactant colorimetric test with a greater than or equal to
100 μg/mL dodecylbenzene sulfonate equivalent cutoff. Jones
and Esposito51 described a modified methylene blue procedure
for the detection and quantitation of surfactants in urine.
Based on the analysis of negative samples, an anionic surfactant
level of 100 μg/mL or greater could be considered adulterated,
but most likely, adulterated specimens will have levels
greater than 800 μg/mL.51
On-Site Adulteration Detection Devices
(Dipsticks) for Urine Specimens
Standard urinalysis test strips such as Multistix (Bayer
Diagnostics, Tarrytown, NY) and Combur-Test (Roche
Diagnostics) are sometimes used to detect the presence of adulterants
in urine. However, among various pads in the test strip,
only pads for detection of nitrite, pH, and specific gravity have
some value. The specific gravity test does not differentiate
between a specific gravity of 1.000 and 1.005 and, therefore, is
very difficult to apply to identify substituted or diluted urine.
The nitrite pad also detects a clinically significant range.
On-site adulterant detection devices have become commercially
available. These dipstick devices offer an advantage
over spot tests because an adulteration check can also be performed
at the collection site. Peace and Tarnai52 evaluated the
performance of 3 on-site devices: Intect 7, MASK Ultrascreen,
and AdultaCheck 4. Intect 7 simultaneously tests for creatinine,
nitrite, glutaraldehyde, pH, specific gravity, PCC, and bleach.
Ultrascreen tests for creatinine, nitrite, pH, specific gravity, and
oxidants. AdultaCheck 4 tests for creatinine, nitrite, glutaraldehyde,
and pH. Peace and Tarnai52 adulterated urine specimens
with Stealth, Urine Luck, Instant Clean ADD-IT-ive, and Klear
at optimum use concentrations and concluded that Intect 7 was
most sensitive and correctly identified adulterants.
AdultaCheck 4 did not detect Stealth, Urine Luck, or Instant
Clean ADD-IT-ive. Ultra Screen detected a broader range of
adulterants than AdultaCheck 4. However, in practice, it only
detected these adulterants at levels well above their optimum
usage, making it less effective than Intect 7.52
However, King53 reported that AdultaCheck 4 is an excellent
way to detect contamination in urine specimens.
AdultaCheck 6 and Intect 7 test strips can determine a range
of creatinine values, although the precise concentration of creatinine
cannot be determined. Similarly, neither test strip can
determine the precise pH of a urine specimen but can only
show the range. However, AdultaCheck 6 and Intect 7 test
strips are capable of successfully identifying urine specimens
with abnormally low creatinine levels and/or pH.54
Adulteration of Hair and Saliva
Specimens for Drug Testing
Hair and saliva are considered alternative specimens to
urine for drugs-of-abuse testing. Drugs can become trapped in
the segment of hair as it develops in a hair follicle. As the hair
segment emerges from the follicle and becomes keratinized, it
carries the drug trapped in the hair. This trapping of drugs permits
analysis a few months after the actual abuse and has a
much longer window of detection than urine or saliva testing.
However, several factors may influence amount of drug
trapped in the hair.
Rollins et al55 studied the effect of hair color on the
amount of codeine trapped in hair follicles by using volunteers
and codeine as a model compound. They observed a strong
correlation between hair concentration of codeine and
melanin concentrations. After 5 days of codeine exposure,
100% of samples from subjects with black hair showed
codeine concentrations higher than the suggested federal
guideline of 200 pg/mg, 50% of subjects with brown hair
demonstrated codeine levels above the cutoff, and subjects
with blonde or red hair showed values less than 200 pg/mg.55
The presence of THC-COOH in oral fluid is a better indicator
of recent use of marijuana than detection of the metabolite
in urine. However, concentrations of drugs in oral fluid are
low, and small amounts of saliva collected present analytic
challenges for saliva testing.56 The main advantage of saliva
and hair testing is that the donor has little chance to adulterate
the specimen. Saliva testing is already in use in the transportation
and insurance industries, and there is increasing interest
in saliva testing in the workplace and for roadside testing for
driving under the influence of drugs. It also seems that given
the advances of technology and reproducibility of test results,
oral fluid test results will be able to survive legal challenges.57
Several products are available for sale through the
Internet with claims that by washing the hair with these products,
a person can pass a drug test. Clear Choice Hair Follicle
Shampoo claims to remove all residues and toxins within 10
minutes of use. One application is sufficient for shoulder
length hair, and the effect can last for 8 hours. Root Clean hair
cleansing system shampoo is also commercially available.
However, no systematic study has been reported of the effects
of using these products to pass a drug test.
Although the chances of adulteration of saliva specimens
are very low to nonexistent, the Saliva Multi-Drug Screen Test
5 (THC, cocaine, opiate, methamphetamine, and benzodiazepine)
is commercially available so that a person can self-test
for the presence of these drugs at home in a saliva specimen
before providing a specimen for actual testing. Oratect multiple
drug screen oral fluid testing is a saliva-based test for several
drugs (cutoffs: amphetamines, 50 ng/mL; benzodiazepine,
20 ng/mL; cocaine, 100 ng/mL; marijuana, 50 ng/mL; opiates,
20 ng/mL; and PCP, 4 ng/mL). A mouthwash is available commercially
(Welcome to I PASSED MY DRUG TEST .COM. Pass any drug test!!) claiming that
rinsing the mouth twice with this product can help a person
pass saliva-based drug testing, which is a popular method of
testing by insurance companies. Another product from the
same company claims to clean hair of any drugs or toxins if the
specially formulated shampoo is used. However, the effectiveness
of such products in escaping detection by drug tests has
not been established by scientific research.
Wong et al58 studied in detail the effects of commercially
available adulterants and foodstuffs on oral fluid drug testing.
An on-site oral fluid drug screen (Oratect) showed no effect of
American Continental or ethnic food (Asian and Hispanic)
on the drug screen result when the adulterant or food was
consumed 30 minutes before analysis. Similarly, common
beverages, including orange and apple juice, did not cause
false-positive results in drugs-of-abuse screening. Cosmetics,
toothpaste, mouthwash, and cigarettes also did not show any
effect. Wong et al58 also evaluated the commercially available
oral fluid adulterants (Clear Choice, Fizzy Flush, and Spit n
Kleen Mouthwash) for their ability to cause a false-negative
drug test result and reported that these products are not capable
of destroying drugs of abuse in saliva specimens.
True-Positive Results in Drugs-of-Abuse
Screening
Ingestion of Poppy Seed Cakes
Poppy seed contains opiates, and consuming poppy seed
cakes or muffins can produce a positive screening result and,
eventually, GC-MS confirmation of morphine and possibly
codeine in urine. The concentrations of morphine and codeine
were 2,797 ng/mL and 214 ng/mL, respectively, in a healthy
volunteer who ingested 3 poppy seed bagels. Opiate was present
in the urine 25 hours after ingestion.59 No opiate was present
45 hours after ingestion. To circumvent this problem, the
Department of Health and Human Services increased the
screening cutoff of opiate immunoassays from 300 ng/mL to
2,000 ng/mL.
In 1 study, 4 volunteers ate 3 poppy seed bagels each.
Neither morphine nor codeine was detected in oral fluids.
However, the levels of morphine ranged from 312 to 602
ng/mL in urine. When 3 volunteers ate 1 poppy seed bagel and
then an unlimited amount of poppy seeds in 1 hour (volunteer
1, 14.82 g; volunteer 2, 9.82 g; and volunteer 3, 20.82 g), the
oral fluid tested positive up to 1 hour after ingestion at a 40-
ng/mL cutoff (highest morphine level, 205 ng/mL). Urine
specimens were positive for 8 hours.60 In Germany, a blood
level of free morphine should be less than 10 ng/mL in drivers.
Moeller et al61 studied blood and urine morphine levels
after subjects consumed poppy seed products. All 5 volunteers
showed positive opiate urine drug test results (up to 2,079
ng/mL by a semiquantitative Abbott assay; urine morphine
level, 147-1,300 ng/mL by GC-MS). No blood specimen tested
positive for free morphine but yielded morphine levels up
to 24 ng/mL following hydrolysis.61
Recently, Hill et al62 studied the effect of ingesting a large
amount of poppy seeds on the urinary concentrations of morphine
in volunteers. The poppy seed study was performed
using Australian poppy seeds because they contain the largest
amount of morphine of any poppy seeds available on the US
market. The morphine content of Australian poppy seeds
ranges from 90 to 200 μg of morphine per gram of poppy seed,
whereas Dutch and Turkish poppy seeds contain only 4 to 5 μg
of morphine per gram of poppy seed. Ten subjects (6 male and
4 female) ingested 2 servings of poppy seed pastry per week
(generally on Monday and Tuesday; 8.1 mg average morphine
per serving) for 3 weeks (total morphine consumed, 49 mg).
Hair specimens were obtained before and after the study.
Urine specimens were obtained for a 24-hour period following
poppy seed ingestion and 3 to 5 hours after subsequent
ingestion of the poppy seed pastry. The maximum values of
urinary morphine ranged from 2,929 to 13,857 ng/mL (as
determined by mass spectrometry). Moreover, urinary morphine
levels remained higher than the 2,000 ng/mL cutoff for
as long as 10 hours. Of 10 subjects, 7 reported drowsiness for
1 hour after eating the poppy seed pastry. The effect lasted 2
to 4 hours. Despite high urinary morphine levels, all subjects
had hair levels of morphine less than the standard cutoff (0.04-
0.48 ng/10 mg of hair; cutoff, 2 ng/10 mg of hair).62
Health Inca Tea and Mate de Coca Tea
Drinking Health Inca Tea may cause a positive test result
for the cocaine metabolite benzoylecgonine. Although US
Customs regulations require that no cocaine be present in any
herbal tea, literature references indicate that some Health Inca
Tea sold in the United States contains cocaine. Jackson et al60
reported a urinary concentration of benzoylecgonine after
ingestion of a cup of Health Inca Tea by volunteers.
Benzoylecgonine was present even 26 hours after ingestion.
The maximum urinary benzoylecgonine concentration of 1.4
to 2.8 mg/L was observed 4 to 11 hours after ingestion of
Health Inca Tea.63
Mate de Coca is a commercially available tea made from
coca leaves (Erythroxylon coca). Turner et al64 prepared tea by
allowing 1 Mate de Coca tea bag to be immersed in 250 mL
of boiling water for 25 minutes. The bag was removed and
squeezed into the tea to drain additional water. A 5-mL sample
was taken for analysis, and volunteers drank the rest. Urine
samples were obtained at 2, 5, 8, 15, 21, 24, 43, and 68 hours
after drinking tea. All urine samples tested positive for benzoylecgonine,
the metabolite of cocaine, by immunoassay.
The amount of cocaine in tea was estimated to be 2.5 mg.64
Passive Inhalation of Marijuana
Exceeding the cutoff limit for marijuana tests is difficult
to achieve through passive inhalation. The Department
of Transportation indicated that MROs should not recognize
passive drug exposure as a legitimate medical explanation
for a positive test result.13,65 THC released in air is
most likely to exist incorporated as an aerosol particle
with a concentration following mixing that becomes highly
diluted.
One study indicated that passive inhalation of marijuana
may lead to positive screening and confirmation test results in
oral fluid up to 30 minutes after exposure.66 Niedbala et al67
later studied the effects of passive inhalation of marijuana on
urine and oral fluid testing results using high marijuana—containing
cigarettes. In study 1, 4 smokers smoked THC mixed
with tobacco (39.5 mg of THC) in an unventilated 8-passenger
van, and 4 volunteers were passive smokers. In study 2, 4
volunteers smoked marijuana only (83.2 mg of THC). Oral
fluid specimens were obtained using the Intercept Oral
Specimen Collection Device (OraSure Technology,
Bethlehem, PA). Participants were allowed to go outside the
van 60 minutes after exposure. Oral fluid specimens were
obtained at baseline (30 minutes before exposure); 0, 15, and
45 minutes inside the van; and 1, 1.25, 1.5, 1.75, 2, 2.5, 3.5,
4, 6, and 8 hours outside the van. Oral fluid specimen collection
continued up to 72 hours after exposure. Urine specimens
were also obtained. Oral fluid was tested for THC
metabolites by using the Cannabinoids Intercept Micro-Plate
enzyme immunoassay with a cutoff of 3 ng/mL (for the confirmation
assay, 2.0 ng/mL). For urine specimens, a 50-
ng/mL cutoff was used. All urine specimens tested negative
(cutoff, 50 ng/mL) for all passive smokers (GC-MS showed
THC metabolite concentrations in the range of 5.8 to 14.7
ng/mL 6 to 8 hours after exposure). In study 1, in which oral
fluid specimens were collected in the van, some subjects
showed a positive response owing to contamination of the
oral fluid collection device with THC smoke, but in study 2,
when all oral fluid specimens were obtained outside the van,
no positive specimens were observed. On the other hand, all
smokers showed significant THC in oral fluid and urine as
expected. The mean urinary concentration was 75 ng/mL 4
hours after smoking.67
Ingestion of Products Containing Hemp Oil
A positive cannabinoid workplace drug test following
ingestion of a commercially available hemp oil preparation
has been reported. The first specimen with a negative test
result was 53 hours after ingestion.68 Alt and Reinhardt69
reported the presence of THC metabolites 80 hours after the
ingestion of 40 to 90 mL of hemp seed oil by volunteers.
THC was also detected in blood samples as early as 1 hour
after ingestion. Alt and Reinhardt69 also reported the presence
of THC in hemp food products such as hemp bar, hemp flour,
and hemp liquor.
On the other hand, Hemp Ale, an alcoholic beverage
formulated, brewed, and bottled by Frederick Brewing
(Frederick, MD) does not contain any THC. The manufacturer
states that the amount of hemp seed added during the
brewing process varies less than 1% between batches. The
hemp seeds go through 2 wash cycles before brewing so that
vegetative material that potentially may contain THC is
removed. Gibson et al70 reported the absence of THC in Hemp
Ale and concluded that ingestion of a moderate amount of the
drink is not sufficient to produce a positive cannabis drug
screen result. Although some studies showed that consumption
of hemp oil can produce a positive THC, Leson et al71
reported that consumption of 125 mL of hemp oil (0.6 mg of
THC) produced a THC metabolite level of only 5.2 ng/mL.
Gustafson et al72 used 7 volunteers in their study (doubleblind,
placebo-controlled) who received 0, 0.39, 0.47, 7.5, and
14.8 mg/d of THC. No specimens were positive for the low
doses (0.39 and 0.47 mg). Only the high doses (7.5 and 14.8
mg) produced a positive THC test result (50- and 100-ng/mL
cutoffs, respectively). However, only 23.2% to 45.8% of specimens
showed positive results.72
Conclusions
Adulterants impose a new challenge in the testing for
abused drugs. Routine specimen integrity testing involving
pH, creatinine, specific gravity, and temperature is inadequate
for detecting the presence of more recently introduced adulterants
such as Urine Luck, Klear, and Stealth. These agents can
cause false-negative results in immunoassay screening steps
and may also affect the GC-MS confirmation step if the
abused drugs are present in modest concentrations (100%-
150% of cutoff concentrations). Therefore, the true presence
of a drug can be missed if these agents are used for adulteration.
Fortunately, spot tests have been introduced, and several
dipstick tests (eg, AdultaCheck 4, AdultaCheck 6, and Intect
7) are available for validation of specimen integrity. Studies
are also needed of the effectiveness of hair shampoo for causing
false-negative results in a hair drug test and mouthwash
products for invalidating saliva testing for abused drugs.
Ingestion of poppy seed cake and hemp oil leads to true-positive
results with opiate and THC assays, and laboratory professionals
should also be familiar with this information.
References
1. Substance Abuse and Mental Health Services Administration,
Department of Health and Human Services. Mandatory
Guidelines for Federal Workplace Drug Testing Programs.
Fed Regist. 1988;53:11970-11980.
2. Substance Abuse and Mental Health Services Administration,
Department of Health and Human Services. Mandatory
Guidelines for Federal Workplace Drug Testing Programs.
Fed Regist. 1994;59:29908-29912.
3. Substance Abuse and Mental Health Services Administration,
Department of Health and Human Services. Mandatory
Guidelines for Federal Workplace Drug Testing Programs.
Fed Regist. 1997;62:51118-51120.
4. Substance Abuse and Mental Health Services Administration,
Department of Health and Human Services. Mandatory
Guidelines for Federal Workplace Drug Testing Programs.
Fed Regist. 2004;19643-19673.
5. Paul B, Jacobs A. Spectrophotometric detection of iodide and
chromic (III) in urine after oxidation to iodine and chromate
(VI). J Anal Toxicol. 2005;29:658-663.
6. Burrows DL, Nicolaides A, Rice PJ, et al. Papain: a novel urine
adulterant. J Anal Toxicol. 2005;29:275-295.
7. George S, Braithwaite RA. An investigation into the extent
of possible dilution of specimens received for urinary drugs of
abuse screening. Addiction. 1995;90:967-970.
8. Needleman SD, Porvaznik M. Creatinine analysis in single
collection urine specimens. J Forensic Sci. 1992;37:1125-1133.
9. Beck O, Bohlin M, Bragd F, et al. Adulteration of urine drug
testing: an exaggerated cause of concern [in Swedish]
Lakartidningen. 2000;97:703-706.
10. Fraser AD, Zamecnik J. Impact of lowering the screening
and confirmation cutoff values for urine drug testing based
on dilution indicators. Ther Drug Monit. 2003;25:723-727.
11. Soldin SJ, Morales AJ, D'Angelo LJ, et al. The importance
of lowering the cut-off concentrations of urine screening and
confirmatory tests for benzoylecgonine/cocaine [abstract]. Clin
Chem. 1991;37:993.
12. Luzzi VI, Saunders AN, Koenig JW, et al. Analytical
performance of immunoassays for drugs of abuse below
established cutoff values. Clin Chem. 2004;50:717-722.
13. Department of Transportation. Part 40: procedures for
transportation workplace drug and alcohol testing. Fed Regist.
2000;65:79462-79579.
14. Edgell K, Caplan YH, Glass LR, et al. The defined HHS/DOT
substituted urine criteria validated through controlled
hydration study. J Anal Toxicol. 2002;26:419-423.
15. Barbanel CS, Winkelman JW, Fischer GA, et al. Confirmation
of the Department of Transportation criteria for a substituted
urine specimen. J Occup Environ Med. 2002;44:407-416.
16. Cook JD, Hannon MW, Vo T, et al. Evaluation of freezing
point depression for classifying random urine specimens
defined as substituted under HHS/DOT criteria. J Anal
Toxicol. 2002;26:424-429.
17. Cone EJ, Lange R, Darwin WD. In vivo adulteration: excess
fluid ingestion causes false negative marijuana and cocaine
urine test results. J Anal Toxicol. 1998;22:460-473.
18. Mikkelsen SL, Ash KO. Adulterants causing false negatives
in illicit drug testing. Clin Chem. 1988;34:2333-2336.
19. McNamara CE, Perry NB, Follett JM, et al. A new glucosyl
feruloyl quinic acid as a potential marker for roots and
rhizomes of goldenseal, Hydrastis canadenis. J Nat Prod.
2004;67:1818-1822.
Deventer K, Delbeke FT, Roels K, et al. Screening for 18
diuretics and probenecid in doping analysis by liquid
chromatography—tandem mass spectrometry. Biomed
Chromatogr. 2002;16:529-535.
21. Warner A. Interference of household chemicals in
immunoassay methods for drugs of abuse. Clin Chem.
1989;35:648-651.
22. Uebel RA, Wium CA. Toxicological screening for drugs of
abuse in samples adulterated with household chemicals. S Afr
Med J. 2002;92:547-549.
23. Schwarzhoff R, Cody JT. The effects of adulterating agents
on FPIA analysis of urine for drugs of abuse. J Anal Toxicol.
1993;17:14-17.
24. Baiker C, Serrano L, Lindner B. Hypochlorite adulteration of
urine causing decreased concentration of delta-9-THCCOOH
by GC/MS. J Anal Toxicol. 1994;18:101-103.
25. Stolk LM, Scheijen JL. Urine adulteration with denture
cleaning tablets [letter]. J Anal Toxicol. 1997;21:403.
26. Pearson SD, Ash KO, Urry FM. Mechanism of false-negative
urine cannabinoid immunoassay screens by Visine eyedrops.
Clin Chem. 1989;35:636-638.
27. Fogerson R, Schoendorfer D, Fay J, et al. Qualitative detection
of opiates in sweat by EIA and GC-MS. J Anal Toxicol.
1997;21:451-458.
28. Rojsitthisak P, Wichitnithad W, Pipitharome O, et al. Simple
HPLC determination of benzalkonium chloride in ophthalmic
formulations containing antazoline and tetrahydrozoline. PDA
J Pharm Sci Technol. 2005;59:323-327.
29. Wu A, Bristol B, Sexton K, et al. Adulteration of urine by
Urine Luck. Clin Chem. 1999;45:1051-1057.
30. Paul BD, Martin KK, Maguilo J Jr, et al. Effects of pyridinium
chlorochromate adulterant (Urine Luck) on testing of drugs of
abuse and a method for quantitative detection of chromium
(VI) in urine. J Anal Toxicol. 2000;24:233-237.
31. Dasgupta A, Wahed A, Wells A. Rapid spot tests for detecting
the presence of adulterants in urine specimens submitted for
drug testing. Am J Clin Pathol. 2002;117:325-329.
32. Ferslew KE, Hagardorn AN, Roberts TA. Capillary ion
electrophoresis of endogenous anions and anionic adulterants
in human urine. J Forensic Sci. 2001;46:615-626.
33. Ferslew KE, Nicolaides AN, Robert TA. Determination of
chromate adulteration of human urine by automated
colorimetric and capillary ion electrophoretic analyses. J Anal
Toxicol. 2003;27:36-39.
34. Paul BD. Six spectroscopic methods for detection of oxidants
in urine: implication in differentiation of normal and
adulterated urine. J Anal Toxicol. 2004;28:599-608.
35. ElSohly MA, Feng S, Kopycki WJ, et al. A procedure to
overcome interferences caused by adulterant "Klear" in the
GC-MS analysis of 11-nor-Δ9-THC-9-COOH. J Anal Toxicol.
1997;20:240-242.
36. Tsai SC, ElSohly MA, Dubrovsky T, et al. Determination of
five abused drugs in nitrite-adulterated urine by immunoassay
and gas chromatography—mass spectrometry. J Anal Toxicol.
1998;22:474-480.
37. Tsai LS, ElSohly MA, Tsai SF, et al. Investigation of nitrite
adulteration on the immunoassay and GC-MS analysis of
cannabinoids in urine specimens. J Anal Toxicol. 2000;24:708-714.
38. Urry F, Komaromy-Hiller G, Staley B, et al. Nitrite
adulteration of workplace drug testing specimens: sources and
associated concentrations of nitrite and distinction between
natural sources and adulteration. J Anal Toxicol. 1998;22:89-
95.
39. Singh J, Elberlind JA, Hemphill DG, et al. The measurement
of nitrite in adulterated urine samples by high performance ion
chromatography. J Anal Toxicol. 1999;23:137-140.
40. Kinkennon AE, Black DL, Robert TA, et al. Analysis of nitrite
in adulterated urine samples by capillary electrophoresis.
J Forensic Sci. 2004;49:1094-1100.
41. Cody JT, Valtier S. Effects of Stealth adulteration on
immunoassay testing for drugs of abuse. J Anal Toxicol.
2001;25:466-470.
42. Cody JT, Valtier S, Kuhlman J. Analysis of morphine and
codeine in samples adulterated with Stealth. J Anal Toxicol.
2001;25:572-575.
43. Valtier S, Cody JT. A procedure for the detection of Stealth
adulterant in urine samples. Clin Lab Sci. 2002;15:111-115.
44. George S, Braithwaite RA. The effect of glutaraldehyde
adulteration of urine specimens on Syva EMIT II drugs of
abuse assays. J Anal Toxicol. 1996;20:195-196.
45. Wu AH, Forte E, Casella G, et al. CEDIA for screening drugs
of abuse in urine and the effect of adulterants. J Forensic Sci.
1995;40:614-618.
46. Goldberger BA, Caplan YH. Effect of glutaraldehyde
(UrinAid) on detection of abused drugs in urine by
immunoassay [letter]. Clin Chem. 1994;40:1605-1606.
47. Wu A, Schmalz J, Bennett W. Identification of Urin-Aid
adulterated urine specimens by fluorometric analysis [letter].
Clin Chem. 1994;40:845-846.
48. Cody JT, Schwarzhoff RH. Impact of adulterants on RIA
analysis of urine for drugs of abuse. J Anal Toxicol.
1989;13:277-284.
49. Lewis SA, Lewis LA, Tuinman A. Potassium nitrite reaction
to 11-nor-delta 9-tetrahydrocannabinol-9-carboxylic acid in
urine in relation to the drug screening analysis. J Forensic Sci.
1999;44:951-955.
50. Paul BD, Jacobs A. Effects of oxidizing adulterants on
detection of 11-nor-Δ9-THC-9-carboxylic acid in urine. J Anal
Toxicol. 2002;26:460-463.
51. Jones JT, Esposito FM. An assay evaluation of the methylene
blue method for the detection of anionic surfactants in urine.
J Anal Toxicol. 2000;24:323-327.
52. Peace MR, Tarnai LD. Performance evaluation of three on-site
adulteration detection devices for urine specimens. J Anal
Toxicol. 2002;26:464-470.
53. King EJ. Performance of AdultaCheck 4 test strips for the
detection of adulteration at the point of collection of urine
specimens used for drugs-of-abuse testing. J Anal Toxicol.
1999;23:72.
54. Dasgupta A, Chughtai O, Hannah C, et al. Comparison of
spot tests with AdultaCheck 6 and Intect 7 urine test strips
for detecting the presence of adulterants in urine specimens.
Clin Chim Acta. 2004;348:19-25.
55. Rollins DE, Wilkins DG, Krueger GG, et al. The effect of hair
color on the incorporation of codeine into human hair. J Anal
Toxicol. 2003;27:545-551.
56. Kintz P, Bernhard W, Villain M, et al. Detection of cannabis
use in drivers with the Drugwipe device and by GC-MS after
intercept device collection. J Anal Toxicol. 2005;29:724-727.
57. Kadehjian L. Legal issues in oral fluid testing. Forensic Sci Int.
2005;150:151-160.
58. Wong RC, Tran M, Tung JK. Oral fluid testing: effects of
adulterants and foodstuffs. Forensic Sci Int. 2005;150:175-180.
59. Struempler RE. Excretion of codeine and morphine following
ingestion of poppy seeds. J Anal Toxicol. 1987;11:97-99.
60. Jackson GF, Saddy JJ, Poklis A. The determination of
morphine in urine and oral fluid following ingestion of poppy
seeds. J Anal Toxicol. 2003;27:449-452.
61. Moeller MR, Hammer K, Engel O. Poppy seed consumption
and toxicological analysis of blood and urine samples. Forensic
Sci Int. 2004;143:183-186.
62. Hill V, Cairns T, Cheng CC, et al. Multiple aspects of hair
analysis for opiates: methodology, clinical and workplace
population, codeine and poppy seed ingestion. J Anal Toxicol.
2005;29:696-703.
63. Rohrig TP, Moore C. Urinary excretion of benzoylecgonine
following ingestion of Health Inca Tea. Forensic Sci Int.
1991;49:57-64.
64. Turner M, McCrory P, Johnston A. Time for tea anyone?
Br J Sports Med. 2005;39:e37.
65. Mule SJ, Lomax P, Gross SJ. Active and realistic passive
marijuana exposure tested by three immunoassays and
GC/MS. J Anal Toxicol. 1988;12:113-116.
66. Niedbala S, Kardos S, Salamone D, et al. Passive cannabis
smoke exposure and oral fluid testing. J Anal Toxicol.
2004;28:546-552.
67. Niedbala RS, Kardos KW, Fritch D, et al. Passive cannabis
smoke exposure and oral fluid testing, II: two studies of
extreme cannabis smoke exposure in a motor vehicle. J Anal
Toxicol. 2005;29:607-615.
68. Struempler RE, Nelson G, Urry TM. A positive cannabinoid
workplace drug test following the ingestion of hemp seed oil.
J Anal Toxicol. 1997;21:283-285.
69. Alt A, Reinhardt G. Positive cannabis results in urine and
blood samples after consumption of hemp food products
[letter]. J Anal Toxicol. 1998;22:80-81.
70. Gibson CR, Williams RD, Browder RO. Analysis of Hempen
Ale for cannabinoids [letter]. J Anal Toxicol. 1998;22:179.
71. Leson G, Pless P, Grotenhermen F, et al. Evaluating the impact
of hemp food consumption on workplace drug tests. J Anal
Toxicol. 2001;25:691-698.
72. Gustafson R, Levine B, Stout P, et al. Urinary cannabinoid
detection times after controlled oral administration of delta-9-
tetrahydrocannabinol to humans. Clin Chem. 2003;49:1114-
1124.


Source: The Effects of Adulterants and Selected Ingested Compounds on Drugs-of-Abuse Testing in Urine
 
Back
Top Bottom