JASBC Extraction Efficiency Dry-Hopping PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/333856224

Extraction Efficiency of Dry-Hopping

Article  in  Journal of the American Society of Brewing Chemists · June 2019


DOI: 10.1080/03610470.2019.1617622

CITATIONS READS

0 102

3 authors:

Dean G. Hauser Scott Lafontaine


Oregon State University University of California, Davis
7 PUBLICATIONS   3 CITATIONS    23 PUBLICATIONS   55 CITATIONS   

SEE PROFILE SEE PROFILE

T.H. Shellhammer
Oregon State University
86 PUBLICATIONS   1,751 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The Efficiency of Dry-hopping and Strategies for the Better Utilization of Dry-hops View project

Hop Enzymes View project

All content following this page was uploaded by Dean G. Hauser on 18 June 2019.

The user has requested enhancement of the downloaded file.


Journal of the American Society of Brewing Chemists
The Science of Beer

ISSN: 0361-0470 (Print) 1943-7854 (Online) Journal homepage: https://www.tandfonline.com/loi/ujbc20

Extraction Efficiency of Dry-Hopping

Dean G. Hauser, Scott R. Lafontaine & Thomas H. Shellhammer

To cite this article: Dean G. Hauser, Scott R. Lafontaine & Thomas H. Shellhammer (2019):
Extraction Efficiency of Dry-Hopping, Journal of the American Society of Brewing Chemists, DOI:
10.1080/03610470.2019.1617622

To link to this article: https://doi.org/10.1080/03610470.2019.1617622

Published online: 18 Jun 2019.

Submit your article to this journal

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ujbc20
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS
https://doi.org/10.1080/03610470.2019.1617622

Extraction Efficiency of Dry-Hopping


Dean G. Hauser , Scott R. Lafontaine , and Thomas H. Shellhammer
Department of Food Science & Technology, Oregon State University, Corvallis, OR, USA

ABSTRACT KEYWORDS
Dry-hopping, which is the cold, aqueous extraction of hops into beer, has gained popularity in Beer; dry-hopping;
recent years as a method for achieving intense hop aroma and flavor in beer. With some brewers efficiency; Humulus lupulus;
dry-hopping at rates up to 2 kg/hL (5 lb/barrel), considerable waste is generated in terms of both spent hops; sustainability
beer loss and spent hops. The retention of valuable volatile and non-volatile hop components
within spent hops, as well as their extraction into beer, was investigated on both a pilot and
industrial scale. On the pilot scale (80 L), an unhopped pale ale was statically dry-hopped with dif-
fering lots of whole cone AmarilloV, Centennial, and Cascade at a rate of 386 g/hL (1 lb/bbl) for
R

24 h at 13  C (55  F). Spent hop material was also collected from a local brewery that dry-hopped
60 hL (50 bbl) of beer at a rate of 1,592 g/hL (4.125 lb/bbl). Approximately one-third of the dry-
matter composition of hops was lost during dry-hopping regardless of hop variety; however, there
was high retention of both a-acids (77% pilot scale, 52% industrial) and hop essential oil (51%
pilot scale, 33% industrial). The oil remaining in the spent hops was enriched in hydrocarbon com-
pounds and depleted in oxygenated compounds. These results indicate that spent dry-hops con-
tain considerable brewing value and have the potential for re-use.

Introduction the form of either volatile or non-volatile hop components.


The non-volatile a-acids are the primary precursors of bit-
Dry-hopping is the cold, aqueous extraction of hops during
terness in beer, and undergo heat-catalyzed isomerization to
or after fermentation.[1] This practice achieves a distinctive
the iso-a-acids during wort boiling.[5] The solubility of
aroma that is notably different from kettle or late-hop addi- a-acids in wort at room temperature and pH 5 has been
tions, likely owing to the better transfer of volatile aroma estimated at 40 mg/L,[6] although reported values in dry-
compounds from hops to beer and/or better retention in hopped beer are substantially lower and vary depending on
beer in a closed system and at lower temperatures.[2] This hopping regime on both the hot and cold side of brew-
technique has become a key tool for American craft brewers ing.[7–9] Iso-a-acids, however, are more soluble in wort and
who produce hop-forward beers, with the output of dry- beer and can be found in finished beer with a published
hopped beer already exceeding one million hectoliters per concentration range of 15–80 mg/L.[6] Brewers’ use of the
year at some breweries.[3] The effect of this trend is echoed term “hop utilization” refers to the quantity of iso-a-acids
by those seen in the American hops industry; between 2010 found in the beer relative to the amount of a-acids added
and 2017 the ratio of aroma to dual-purpose hop acreage during wort boiling, and typically ranges from 10 to 40%
shifted from roughly 50/50 to 80/20.[4] Dry-hopping rates depending on factors such as the form of the hop (cone,
among American craft brewers can range well above 2 kg/hL pellet, extract), timing of addition, and physical properties
(5lb/bbl), constituting a hop dose that is often one or two of the wort.[10] Studies on the kinetics of a-acid isomeriza-
orders of magnitude larger than a typical kettle addition for tion have shown that even at 80  C for 90 min, isomerization
lager beer, making dry-hopping on an industrial scale yield can be well below 10%.[11] These data provide a rea-
expensive and potentially unsustainable. In addition to the sonable basis to assume that at the lower temperatures of
increased dosage, aroma varieties used for dry-hopping often dry-hopping (typically below room temperature), the solv-
command a higher price, and the practice of dry-hopping ation and isomerization of a-acids would be minimal and,
can also lead to significant beer loss during the separation of therefore, suggest that a significant proportion of them may
hop material from dry-hopped beer. In short, dry-hopping remain in the spent hops.
at high rates can be expensive and has a potentially negative Hop aroma in beer is primarily driven by the volatile oils
impact on the profitability of these types of beers. found within the hop cone, which can account for roughly
In addition to beer loss, dry-hopping generates a signifi- 0.5–3% of the dry mass of hops.[12] The major groups
cant amount of waste in terms of the spent hops that are of compounds found within hop oil are hydrocarbons
discarded, which potentially still contain brewing value in (mainly terpenes and aliphatic hydrocarbons), oxygenated

CONTACT Thomas H. Shellhammer tom.shellhammer@oregonstate.edu


Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/ujbc.
ß 2019 American Society of Brewing Chemists, Inc.
2 D. G. HAUSER ET AL.

hydrocarbons (mainly terpenoids), and sulfur compounds Table 1. Hop acid and total oil content of samples used for pilot-scale
dry-hopping.
(thioesters, sulfides, etc.), which account for roughly two-
thirds, one-third, and 1% of the mass of hop oil, respect- Total oil
a-acid (%wt/wt) b-acid (%wt/wt) (mL/100g)
ively.[13] The contributions of specific hop volatiles to hop
aroma and flavor in beer are not well known due to the fact Hop Variety Range Mean Range Mean Range Mean
AmarilloV
R

that over 1,000 individual constituents likely exist within 8.20–9.44 8.88 5.29–6.76 6.18 2.14–2.53 2.34
Cascade 4.51–7.81 5.88 3.55–7.43 5.06 0.80–1.95 1.31
hop oil (of which over 440 have been identified),[14] as well Centennial 8.65–10.93 10.00 2.43–3.81 3.01 1.70–3.56 2.74
as synergistic and masking effects of the combinations of Values represent means and ranges of 8–12 replicates and were calculated on
these volatiles.[15] In spite of these difficulties, a number of a moisture-free basis.
high-impact hop aroma compounds have been identified in
beer, including the terpene alcohols linalool and gera- aroma in dry-hopped beer on hop harvest maturity and
niol.[15,16] There is a large range of reported transfer rates of kilning.[25,26] The hops used in the pilot-scale treatments
these compounds from hops to beer, with values for linalool included the varieties Cascade, Centennial, and AmarilloV
R

and geraniol ranging from 20 to 110%, and 50 to 240%, (Table 1). Shortly after the 2017 harvest, whole cone
respectively, relative to their initial dosage and depending Cascade and Centennial hops were obtained from S. S.
on the parameters of the dry-hopping system.[9,17–19] Many
Steiner, Inc. (Yakima, WA, U.S.A.). Whole cone AmarilloV
R

of the other terpene-based hop oil compounds have low


hops were obtained from Virgil Gamache Farms
reported extraction rates (<5%) in beer due to their non-
(Toppenish, WA, U.S.A.). Upon arrival at Oregon State
polar character.[19] Some of these compounds are lesser con-
University, all hops samples were repackaged in high barrier
tributors to hoppy aroma in beer, for example b-myrcene,
foil pouches, purged with nitrogen, vacuum sealed, and
which by itself may not be important but serves as a precur-
stored cold (–20  C) until dry-hopping and chemical ana-
sor for more high impact odorants such as linalool and
geraniol via autooxidation.[20] It is also worth noting that lysis. On an industrial scale, spent dry-hop slurry was col-
dry-hopping is far from a standardized practice, and the lected from a local brewery in addition to samples of hop
extraction and retention of hop components depend on a pellets that were used for the particular dry-hops addition.
The hops consisted of five varieties (CitraV, AmarilloV,
R R

vast number of system parameters including but not lim-


Centennial, MosaicV, and CTZV), which were added in two
R R

ited to the temperature, hop variety (or varieties), dosage


rate of hop material, hop product being used, addition additions over the course of five days, for a total dosage of
method (direct dispersion or restrained in mesh bags or 1592 g/hL (4.125 lb/bbl). Over two months, four representa-
otherwise), use of mechanical agitation during dry-hopping, tive spent-hops samples were collected from separate pro-
contact time, presence of yeast, and scale of the brewing duction batches of the same brand of beer, brought to
operation.[1,3,7,9,19,21,22] Oregon State University, dewatered using a mesh bag, and
To the best of our knowledge, there is currently no pub- placed in high-barrier foil pouches that were subsequently
lished work on the non-volatile and volatile hop compo- purged with nitrogen, vacuum sealed, and stored at –20  C
nents that are retained within spent dry-hops, in spite of the until analysis. This slurry resulted from the production of
growing body of work looking into transfer rates of various an imperial India pale ale, which had a starting extract con-
hops components into beer. Given the environmental chal- centration of 17.3 Plato, alcohol content of 8.4% (v/v), real
lenges that brewers must contend with (energy and water extract of 4.92 Plato, color of 7.9 SRM, and pH of 4.76.
use plus waste generation) and the evidence that consumers Kettle hop additions were added at a rate of roughly 656 g/
would be willing to pay more for more sustainably brewed hL (1.7 lb/bbl) resulting in an 80 bitterness unit (BU) beer.
beer, re-use of dry-hops waste could potentially provide a
way to add value to a waste stream, while easing the envir-
onmental impacts of beer production.[23,24] With this appli- Beer production
cation in mind, the objectives of this study were to (1) In order to simulate a yeast-free dry-hopping system, an
provide a survey of what brewing value is left within spent unhopped beer was prepared by a commercial brewery in
dry-hops from a non-volatile and volatile perspective on a Portland, Oregon (Bridgeport Brewing). Unhopped wort
pilot and industrial scale, (2) evaluate the fate of individual (41 hL) was prepared with 86.2% pale two row, 13.3%
hop volatiles during pilot-scale dry-hopping by determining
Caramel 10 L and 0.5% Caramel 120 L malt (Great
their extraction rates into beer and retention rates in spent
Western, Vancouver, WA, U.S.A.). The starting extract con-
hops, and (3) determine the effect of hop variety and ana-
centration in the unhopped wort was 11.1 Plato.
lyte-specific dosage on extraction and retention rates of
Fermentation was carried out with BridgePort Brewing
hop components.
Company’s house yeast strain at 19–20  C. Following fer-
mentation and post-clarification, iso-a-acids (IsoHop, John I
Experimental Haas, Yakima, WA, U.S.A.) were added at a target concen-
tration of 18 mg/L. This resulted in a 19 BU, 4.5% alcohol by
Hops collection
volume (%ABV) unhopped base beer. Beer was carbonated
The lab-scale dry-hopping experiments were conducted in and packaged into 60-L stainless kegs, shipped to Oregon
tandem with an investigation into the dependence of hop State University, and held at 2  C until dry-hopping.
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS 3

Pilot-scale dry-hopping hand. The high-barrier pouches were subsequently purged


with nitrogen, vacuum sealed, and stored at –20  C
The pilot-scale dry-hopping process established by Vollmer
until analysis.
and Shellhammer[27] has been shown to be reproducible and
to reduce variation between treatments on a pilot scale. The
method is as follows: 24 h prior to hop addition, the Hop analysis
unhopped beer was removed from the cooler at 4  C and
allowed to warm to approximately 15  C. For each treat- Moisture and soluble solids
ment, 40 L of beer was transferred into two modified 60-L Moisture content was determined by ASBC method Hops-
stainless kegs with a 400 stainless steel opening fitted with a 4A, which is a routine air oven method.[28] However, drying
standard Sankey D-system coupler and modified spear for 1 h at 103  C was found to be insufficient for the spent
(Sabco, Toledo, OH, U.S.A.). The hop treatments were hops samples due to their high moisture content, which
added at rate of 386 g hop/hL unhopped beer (1 lb hop/bar- averaged 76% wt/wt (wet basis). Drying trials revealed that
rel). Whole cone hops were ground into a hop grist using beyond 3 h of drying, negligible changes in mass were
an Avantco 1.5 horsepower MG22 #12 Meat Grinder. All observed (0.01%) and 2 h was used as a standard drying
grinder parts were thoroughly scrubbed with an Aclonox time for wet spent hops.
(White Plains, NY, U.S.A.) solution, washed with cold water, Loss of soluble solids from hop material during dry-hop-
and dried between hop samples, in order to minimize the ping was also considered and found to be an important
generation of heat, and eliminate carry-over of hop material. piece of data for comparing pre- and post-dry-hopping com-
Ground hops were placed into two mesh bags (EcoBag, position of the hops used. Laboratory scale dry-hopping
Ossining, NY, U.S.A.), along with stainless-steel fittings to experiments in an acetate buffer system (buffer method)
ensure the hops sank to the bottom of the keg. The bags were used to verify the percent extraction of soluble solids
were stored inside high barrier pouches flushed with N2 observed in the pilot-scale (beer method) dry-hopping treat-
until the dry-hopping event. Approximately 150 g of each ments. In brief, the laboratory scale dry-hopping treatments
hop sample was set aside for chemical analysis, which was were performed by adding ground hops to 250 mL beer buf-
carried out concurrent with the pilot-scale dry-hopping. For fer (pH 4.3, 0.1 N acetate buffered 5% ethanol) at the same
each dry-hop treatment, the two kegs were temporarily de- hops/diluent ratio as in the pilot-scale dry-hopping (3.86 g/L),
pressurized and opened under a stream of low-pressure and removing them after 24 h by filtration through 113 V
CO2. Simultaneously, the high barrier pouch was opened qualitative filter papers (GE Healthcare, Buckinghamshire,
and the mesh bag containing ground hop grist was added to U.K.). Density and alcohol measurements were carried out
the beer. After the addition, the headspace was flushed with using the Anton Paar Beer Alcolyzer. Real extract (RE) was
CO2 and purged. After purging, the kegs were inverted three then calculated based on the measured alcohol content and
times to ensure proper mixing. During dry-hopping, treat- density of the resulting solutions using Tabarie’s formula,
ments were kept under 10 psi CO2, and the average tem- which deconstructs the specific gravity (SG) of a beer into the
perature of the dry-hopping events ranged from 13.3 SG of the alcohol and the SG of the RE.[29]All alcohol by vol-
to 15  C. ume (%ABV) to SG conversions were carried out using
After 24 h of dry-hopping, the beer was filtered in order ASBC Table 2,[28] and all conversions of SGRE to degrees
to stop the dry-hopping process. Dry-hopping was stopped Plato (%wt/wt) were calculated using a fourth order polyno-
after 24 h because prior work by Wolfe[7] showed that the mial that was accurate to three significant figures.[28]
extraction of key hop volatiles occurred within 24 h during Percentage of soluble solids was determined by calculating
dry-hopping. During filtration the two kegs were blended the ratio between the change in real extract due to dry-hop-
via a three-way fitting prior to entering a plate and frame ping and the amount of dry hop matter entering the system
filter using diatomaceous earth impregnated cellulose pads on a per-liter basis.
(HS2000, Pall Corporation, Port Washington, NY, U.S.A.).
Dissolved oxygen (DO) was monitored during filtration
using an Orbisphere 3100 Portable Oxygen Analyzer (Hach, Non-volatiles
Loveland, CO, U.S.A.). Bright beer was not collected until Reagents and standards. HPLC-grade methanol was
DO was below 110 mg/L. After DO was within specification, obtained from VWR International, BDH analytical (West
bright, filtered beer was collected in a closed 19.6 L (1/6 US Chester, PA, U.S.A.). Phosphoric acid was obtained from
bbl) stainless steel keg with sufficient backpressure to reduce Avantor Performance Materials (Center Valley, PA, U.S.A.).
foaming. Between each filter run, filter pads were changed Ethyl ether and hydrochloric acid were obtained from EMD
to prevent hops and beer carry-over from one treatment to Millipore Corporation (Billerica, MA, U.S.A.). International
the next. Filtered beer was stored at 2  C and under CO2 calibration extract ICE-3 standard was obtained from ASBC.
overpressure (11–12 psi) until chemical analysis.
Spent hops were collected immediately after filtration by HPLC analysis. The concentrations of hop acids in hops
removing the mesh bags from the two treatment kegs, gently and beer samples were determined using ASBC methods
hand-wringing them to remove excess beer (resulting in Hops-14 and Beer-23E under modified HPLC conditions
80% moisture content on a wet basis), depositing them on an Agilent 1200 HPLC[28] (Agilent Technologies,
into a high barrier pouch, and thoroughly mixing them by Waldbronn, Germany). Prior to analysis beer was degassed
4 D. G. HAUSER ET AL.

via filtration through GHP AcrodiscV 13 mm, Pall b-caryophyllene, b-farnesene: 5000, 10000). Target analytes
R

Corporation (East Hills, NY, U.S.A.) syringe filters. Analysis were quantified using the following ions for each analyte:
was performed using a 2.6 mm EVO C-18 100 Å 100 x m/z 55 (2-octanol), m/z 68 (limonene), m/z 69 (b-farnesene,
4.6 mm LC 143 column (Phenomenex, Torrance, CA, geraniol, geranial, nerol, neral, geranic acid, and geranyl
U.S.A.) held at 40  C. A 7-lL aliquot of each sample was acetate), m/z 71 (linalool), m/z 79 (caryophyllene oxide),
injected and the elution was carried out using a flow rate of and m/z 93 (a-pinene, b-pinene, b-myrcene, b-caryophyl-
1.6 mL/min. The solvent gradient was as follows: 10% solv- lene, and a-humulene). The target analyte concentrations in
ent A (reagent water): 90% solvent C (75% MeOH, 24.5% hop oil were then standardized on a per-mass basis using
H2O, 0.5% H3PO4) held for 5 min, then changed to 100% the total oil content determined during hydrodistillation.
solvent D (100% MeOH) over 5 min and held for 2 min,
then returned to 10% solvent A: 90% solvent C over 2 min, Hop volatiles in beer. Headspace-Solid Phase Micro
for a total run time of 14 min. The a- and b-acids were Extraction (HS-SPME) was performed on the dry-hopped
measured at 314 nm. No statistically significant difference beer treatments using a 2 cm 24-gauge divinylbenzene/car-
was found between the results of HPLC analysis on hops boxen/polydimethylsiloxane (DVB/CAR/PDMS) Stableflex
and hops adjusted to 80% moisture with the experimen- fiber with 30/50 lm coating thickness (Supelco, Bellefonte,
tal beer. PA, U.S.A.). Eight mL of each sample was placed into a
20 mL screw top amber vial with 3 g sodium chloride EMD
Millipore (Billerisa, MA, U.S.A.). As an internal standard, 2-
Volatiles
octanol (911 ppb) was added to each vial. A MultiPurpose
Total oil. Total hop essential oil (mL/100g) was determined auto sampler (MPS2; Gerstel, M€ ulheim, Germany) was used
by hydrodistillation according to ASBC method Hops 13.[28] for pre-incubation, stirring, extraction, and injection.
Hop oil was stored in 2.5-mL amber vials with foil lined clo- Samples were pre-incubated for 15 min at 30  C and
sures, and was purged with nitrogen before being stored at adsorbed by piercing the vial septa and exposing the fiber to
–20  C until further analysis. the headspace for 45 min with agitation. After adsorption,
the fiber was desorbed into the GC sample inlet (splitless
Reagents and standards. b-myrcene, (þ)-a-pinene, mode, 250  C) for 10 min. The analytical column was a 30 m
(þ)-b-pinene, linalool, geraniol, citral 1&2, (R)-(þ)-limon- x 250 um x 0.25 um Zebron ZB- 1 MS (Phenomenex,
ene, geranyl acetate, geranic acid, nerol, b-caryophyllene, Torrance, CA, U.S.A.) and ultra-pure helium was used as
a-humulene, b-farnesene, and (-)-caryophyllene oxide were the carrier gas (at constant pressure, 11 psi). The following
obtained from Sigma-Aldrich (St. Louis, MO, U.S.A.). The temperature program was used: 50  C hold for 1 min,
2-octanol was obtained through Alfa Asear (Haverhill, MA, 50–250  C (5  C/min) hold for 11 min, and 250  C hold for
U.S.A.). Hexanes were obtained from J.T. Baker (Center 5 min. The auxiliary line and mass spectrometer were oper-
Valley, PA, U.S.A.). ated at 280 and 180  C, respectively. The mass spectrometer
was operated using electron-impact mode at 70 eV and in
Hop oil constituents. Hop oil compositional analysis was full scan mode set up to detect ions with a mass-to-charge
performed using previously published methodology[30] using ratio (m/z) of 30–350. For all target analytes, 5-point cali-
an HP 6890 gas chromatograph with an Agilent 5972a mass bration curves (11, 22, 44, 89, and 300 ppb) were created.
spectrometer (GC-MS) under modified conditions from Calibration curves were made in a model beer solution (pH
ASBC Hops-17.[26] In brief, a 1% 2-octanol (8026 ppm) solu- 4.3, 0.1 N acetate buffered 5% ethanol) and were prepared
tion was prepared in reagent grade hexane. Hop oils were using the methodology previously described.
diluted to 10% with the 1% 2-octanol/hexane solution in
crimped glass vials. The diluted hop oil (1 mL) was directly
injected into the injection port held at 200  C and operating Statistical analysis
in split mode (1:20) using the septum purge option. The One-way analysis of variance (ANOVA) with multiple com-
analytical column was a 30 m x 250 um x 0.25 um Zebron parisons (Tukey-Kramer HSD) (a ¼ 0.05), two-tailed paired
ZB- 1 MS (Phenomenex, Torrance, CA, U.S.A.) and ultra- t tests (a ¼ 0.05), and simple linear regression analyses were
pure helium was used as the carrier gas (at constant flow carried out using JMP Pro 12 (Buckinghamshire, U.K.).
rate, 1.4 mL/min). The following temperature program was
used: 50  C hold for 1 min, 50–180  C (2  C/min) hold for
10 min, 180–200  C (3  C/min) and 250  C hold for 5 min. Results and discussion
The auxiliary line and mass spectrometer were operated at
Accounting for soluble solids loss in dry-hopping
280 and 180  C, respectively. The mass spectrometer was
operated using electron-impact mode at 70 eV and in full In addition to the compounds studied in this work, hops
scan mode set up to detect ions with a mass-to-charge ratio contain significant quantities of potentially soluble protein,
(m/z) of 30–350. For all target analytes, 4-point calibration ash, and phenolic compounds.[6] In order to account for
curves (100, 200, 300, and 900 ppm) were created. For high these losses, the extent of soluble solids extraction during
concentration target analytes, additional calibration points dry-hopping from the hop material used for this study was
were added (b-myrcene, a-humulene: 10000, 25000, 35000; evaluated by two methods (in beer and “beer buffer”) as
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS 5

Table 2. Comparison of methods for determining soluble hop solids. one-way ANOVA (P ¼ 0.0799). The dependence of the
Extractable solids (% wt/wt, dry basis) retention of a-acids in the spent hops on their initial con-
Hop variety In beer In buffer centration in the hop material was examined by fitting a lin-
AmarilloV 31.3 ± 4.9a 31.6 ± 0.8a ear regression for both the beer and buffer adjustments. In
R

Cascade 38.2 ± 2.6b 34.8 ± 1.6b the beer model, there was no evidence of a correlation (R2
Centennial 33.7 ± 4.8a 33.2 ± 1.2b
¼ 0.027, P ¼ 0.4041). Using the buffer model, there was evi-
Values represent means of 6–12 replicates ± one standard deviation. Means
with same superscripts are not significantly different at P < 0.05 based on dence of a negative correlation between the initial concen-
Tukey’s honestly significant difference (HSD). tration and the retention of a-acids (p ¼ 0.039) but the
linear model did not account for much of the variation in
described earlier (Table 2). A paired t test (p ¼ 0.306) the data (R2 ¼ 0.154) suggesting that more work needs to
revealed no significant difference between the beer and buf- be performed to understand this relationship.
fer methods for calculating soluble hop solids, which aver- Neither a- nor b-acids were detected in any of the experi-
aged 34 and 33% (%wt/wt dry basis) respectively. However, mentally dry-hopped beers. It is expected that very low lev-
there was considerably more within-group variation in the els of these components could be transferred into the beer
beer method. ANOVA (beer method P ¼ 0.0029, buffer in a quiescent dry-hopping system, since previous work in
method P < 0.0001) with post-hoc Tukey-Kramer’s HSD similar dry-hopping systems suggests low level extraction of
revealed variety dependent soluble solids groupings, with a-acids (in the range of a few parts-per-million) is possible
Cascade having significantly higher soluble solids at 38% at 1 lb/bbl dosage.[7,9] The b-acids are generally regarded as
(beer) and 35% (buffer) compared with AmarilloV at 31%
R

insoluble in beer,[13] although recent work has shown that


(beer) and 32% (buffer). Due to the larger variation inherent they may be present in higher concentration in highly
in the beer method, Centennial was grouped with AmarilloV
R

hopped, hazy beers.[34] However, in the case of hazy beers,


in the beer method and with Cascade in the buffer method. the high residual hop acids are likely present as a suspension
These data show that within these varieties nearly one third of insoluble compounds as opposed to completely solvated.
of the original hop mass was transferred into beer during In the study presented herein, all beers were filtered to
dry-hopping. This is a new and surprising finding in indus- arrest the dry-hopping process thereby removing insol-
trial dry-hopping, as most brewers consider hop material to
uble components.
be largely inert, and its contribution to beer as dry-hops is The transfer of hop volatiles from the hop material to
viewed only in terms of transfer of flavor-active compounds.
beer is the primary goal of dry-hopping in general.
For the remainder of this study, unless otherwise noted, val-
Therefore, the fate of the oil fraction of the hop is of par-
ues that have been adjusted for the loss of soluble hop solids
ticular interest when examining the efficiency of dry-hop-
were calculated using the “beer” adjustment.
ping. The essential oil content of the hop material used in
this study was compared before and after dry-hopping as
Bitter acid and total oil content of spent hops determined by hydrodistillation (Figure 1). Retention rates
of hop oil were calculated on a per-mass basis (again
The hops used in this experiment showed a wide range of
accounting for moisture and loss of soluble solids) (Figure
both a-acid and total oil content, within and between the
1, Table 3).
different varieties (Table 1), indicating that pre- (harvest
Considering the hop material leaving the dry-hopping
maturity) and post-processing (kilning) factors both play a
system on a dry solids basis, an average of 76% of hop oil
role in determining the chemical composition of hops,
was recovered with suggestive evidence of a variety effect
which has been demonstrated by previous studies.[31–33] For
(P ¼ 0.0649). However, simply using the values from the
each pilot-scale dry-hopping experiment, the composition of
spent material without adjusting for the loss of hop solids
the hops in terms of a-acids and total oil was compared
pre- and post-dry-hopping (Figure 1). The retention rates of does not account for the fact that roughly one-third of the
these components were calculated based on hop composition dry solids are washed out during dry-hopping. Accounting
pre- and post-dry-hopping on a per-mass basis (accounting for the loss of solids reveals a soluble solids-based variety
for moisture and the loss of soluble hop solids of each indi- effect (P ¼ 0.0047), with significant differences in retention
rates between AmarilloV (56%) and Cascade (42%), and
R

vidual treatment) (Figure 1, Table 3).


Centennial (51%) grouped with both AmarilloV and
R

As expected based on the hydrophobic character of both


the a- and b-acids and their relatively low solubility in water Cascade. The relationship between hop oil dose and the
and beer, these components were found to remain in signifi- extent of retention of hop oil in the spent material was eval-
cant quantities within the spent hop material. On average, uated by fitting a linear regression. A statistically significant
82, 74, and 72% of a-acids were retained within the spent positive correlation was found (P ¼ 0.016), but only 20% of
material for AmarilloV, Cascade, and Centennial respect- the variation in the data was accounted for, indicating the
R

ively, resulting in an overall mean of 77% across all three. need for more information in elucidating the mechanisms
Similarly, b-acids were retained at a rate of 84, 80, and 77% controlling the extraction of hop oil. Although total oil con-
within the spent AmarilloV, Cascade, and Centennial hops
R
tent of hops has been shown to be a poor predictor of
respectively, for an overall average of 81%. There was sug- aroma potential in beer,[35] hoppy aroma in beer is still
gestive evidence of a variety effect on a-acid retention in a driven by components of the oil fraction, indicating that
6 D. G. HAUSER ET AL.

Figure 1. Comparison of total oil (red) and a-acid content (blue) of hops pre- versus post-dry-hopping (top) and retention rates as a function of initial concentration
(bottom) for Cascade (black), AmarilloV (gray), and Centennial (white) hops.
R

Table 3. Comparison of hop acid and total oil retention rates in spent hops.
a-Acid retained (%wt/wt) b-Acid retained (%wt/wt) Oil retained (%v/v)
Hop Variety Range Mean Range Mean Range Mean
AmarilloV 82a 84a 57a
R
65–98 66–99 35–68
Cascade 62–96 74a 71–103 80a 30–60 41b
Centennial 58–80 72a 63–106 77a 39–65 51ab
Values represent means and ranges of 8–12 replicates. Means with same superscripts are not significantly different at P < 0.05 based on Tukey’s honestly signifi-
cant difference (HSD).

based on these data (50% retention) spent dry-hops could retention.[34] Compared with the pilot-scale, static dry-hop-
be a potential source of hop aroma. ping treatments, these values were in general roughly 20%
lower, likely owing to differences in key aspects of how dry-
hopping was carried out, which in this case consisted of
Bitter acid and hop essential oil retention in industrial adding hops directly to beer (rather than in mesh bags),
scale dry-hopping extending the contact time to several days (as opposed to
Industry scale dry-hopping samples were collected and sub- 1 day), and stirring the dry-hopped beer several times.
sequently analyzed for bitter acids and total oil, along with a
pre-dry-hopping control made up of a mixture of the five
Transfer and retention of hop volatiles during pilot-
varieties used (Table 4). These data show a better extraction
scale dry-hopping
of hop components as compared with the pilot-scale dry-
hopping with spent material retaining an average of 47 and In order to gain a deeper understanding of the extraction
33% of total oil on a dry “as-is” and dry, solids adjusted and retention rates of various hop aroma compounds, the
basis, respectively. quantities of specific analytes were measured in the raw
In terms of a-acids, the spent material contained an aver- hops, dry-hopped beer, and spent hops (Table 5). Due to
age of 75 and 52% on a dry “as-is” and dry, solids adjusted the different quantities of total oil and aroma compounds
basis respectively. The b-acids were retained at a consider- that resulted from different harvest times or kilning condi-
ably lower rate than in the pilot-scale trials, with less than tions, the resulting data show a wide range of dosing condi-
20% retention on a dry “as-is” basis. Recent work has shown tions in spite of all dry-hops additions being performed at
that hazy IPAs may have increased concentration, and there- the same mass dosage rate.
fore extraction of b-acids, which could offer one potential Although the quantities of these compounds in spent
explanation for the significant decrease observed in b-acid hops need to be adjusted to reflect the loss of soluble solids
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS 7

Table 4. Comparison of hops before and after dry-hopping at a local brewery.


Sample type Moisturea (% wt/wt) a-Acidsb (% wt/wt) b-Acidsb (% wt/wt) Total oilb (mL/100g)
Hop pellets 9.03 11.55 3.99 1.79
Spent hops “as-is” 75.8 ± 2.7 8.67 ± 1.72 0.75 ± 0.009 0.84 ± 0.13
Values represent a single measurement for hop pellets and means of four replicates of spent hops gathered from different production
batches ± one standard deviation.
a
Expressed as wet-basis.
b
Expressed on a moisture-free basis.

during dry-hopping, it is useful to compare the raw hops Table 5. Content of selected aroma compounds in hops, spent hops,
and spent hops “as is” from a practical perspective, since a and beer.
brewer potentially re-using this material would need data on Raw hops Spent hops (as-is) Beer
the actual spent hop material to be used for reuse in order Compound mg/100g n mg/100g n ug/L n
to calculate dosage for bittering or aroma additions. Using Linalool 2.8–20.7 28 0.2–9.5 28 80–281 28
an “as is” comparison also reveals striking differences Geraniol 1.6–37.5 28 0.5–3.8 24 39–357 28
Caryophyllene oxide 2.7–14.4 28 0.7–6.9 28 n.m.a
between individual compounds in terms of their retention in Geranial 0.6–4.6 28 0.9–1.5 5 24–26 2
spent hop material or transfer to beer. Geranyl acetate 1.5–31.4 28 1.0–18.9 28 21–28 27
The terpene and sesquiterpene hydrocarbons (a- and Neral 0.6–4.1 24 0.4–2.0 8 19–26 2
Geranic acid 2.9–105.4 28 1.0–86.1 28 27–110 28
b-pinene, b-myrcene, limonene, b-caryophyllene, b-farnesene, Nerol 0.9–6.8 28 0.5–3.4 28 34–77 28
and a-humulene) were found at maximal concentrations in the b-Farnesene 26.6–147.0 20 14.4–131.2 20 21–25 10
spent material that was often within a few mg/100g of their b-Myrcene 371.5–2382.1 28 258.3–2163.4 28 14–25 28
a-Pinene 0.7–5.9 28 0.4–5.2 28 n.m.
original concentration. This is interesting considering that the b-Pinene 5.2–37.5 28 3.3–34.6 28 n.m.
spent material only contained an average of 77% of the quan- Limonene 1.6–10.8 28 1.1–10.3 28 n.m.
tity of oil that it entered with. Conversely, oxygenated com- a-Humulene 122.4–565.6 28 98.8–544.0 28 24–28 28
b-Caryophyllene 39.0–181.7 28 32.4–184.7 28 21–23 5
pounds such as the terpene alcohols and aldehydes (linalool,
Values represent concentration ranges of n observations, spent hops data are
geraniol, nerol, geranial, and neral) were found in far lower not adjusted for soluble solids. Volatiles in hops are expressed on a mois-
quantities in the spent material, with maximal concentrations a
ture-free basis.
reaching as low as 10% of the initial concentration. n.m.¼not measured.
The trends in retention were mirrored by the quantities
of these constituents that were transferred to the beer sesquiterpenes.[22] Aqueous solubility is strongly correlated
matrix; terpene hydrocarbons were found in far lower quan- with Kow and also tracks relatively well with the data col-
tities than oxygenated terpenes. Terpene acids, esters, and lected in this study (Table 6). Due to differences in reported
oxides (geranic acid, geranyl acetate, and caryophyllene Kow and solubility data depending on method, selected data
oxide) showed transfer and retention trends intermediate to are reported or calculated from a variety of sources.[37–42]
the terpene hydrocarbons and the terpene alcohols and alde- Transfer and retention rates were calculated based on the
hydes. These trends seem to be dictated by the physical proportion of individual analytes that were found in the
properties of the compounds, namely their polarity, which dry-hopped beer and spent hops relative to the amount of
will be discussed in detail in the next section. analyte that was dosed into the beer. All calculations involv-
ing hop material were carried out using the dry mass of the
Octanol-water partition coefficients as a model for material. Retention rate calculations in spent hops were add-
extraction and retention rates of hop aroma compounds itionally adjusted to account for solids loss. Extraction and
retention rates of the compounds investigated in this study
Octanol-water partition coefficients describe the relative ten- were plotted against log(Kow) in order to provide a visual
dency of a given compound to partition into an organic (octa- representation of the relationship between these parameters
nol) or aqueous phase and are used widely in environmental (Figure 2). These data reinforce the utility of Kow as a model
chemistry in order to trace the fate of organic pollutants.[36] for the fate of hop volatiles during dry-hopping, with clear
These coefficients are also used to assess the distribution of distinctions in the extraction and retention rates for differ-
substances in biological systems[37] and have applications in ent groups of compounds. The sesquiterpene hydrocarbons
toxicology and drug design.[38] The log of the octanol-water (a-humulene, b-farnesene, b-caryophyllene, log(Kow) ¼
partition coefficient (log(Kow)) is the most commonly reported 6.3–7.1) showed some of the lowest mean extraction rates
value, and higher and lower values for log(Kow) designate (0.2–0.6%) and highest mean retention rates (52–63%). This
hydrophobic or hydrophilic character respectively. is particularly interesting given recent work that has rein-
In the context of hop aroma chemistry, Kow has been forced the belief that the oxygenated forms (sesquiterpene
used to design stir bar-sorptive extraction methods for the alcohols and epoxides) of a-humulene and b-caryophyllene
analysis of hop volatiles in beer.[39] Recent work by are likely important for kettle hop aroma.[43] In a similar
Haslbeck et al. introduced the Kow as a model for the trans- manner to the sesquiterpenes, the monoterpene hydrocarbon
fer of hop volatiles into beer, and found good correlation b-myrcene (log(Kow) ¼ 4.3) had the lowest measured mean
between these values and the transfer rates of a range of extraction rate (0.05%), which is likely due to a combination
hop derived C8 and C11-esters, monoterpenes, and of its nonpolar character and the fact that it the most
8 D. G. HAUSER ET AL.

Table 6. Physical properties and extractability of selected hop aroma compounds.


Log (Kow)a (log(mol/ Aqueous solubilityb Extraction from hops into Retention rate in spent Unaccountable processing
Compound mol) 25  C) (ppm, 25  C) beer (% wt/wt) hops (% wt/wt) loss (% wt/wt)
Linalool 3.4–3.5 1542 46 16 38
Geraniol 3.5–3.6 404–877 43 24 33
Geranyl acetate 4.0 18 9 36 55
b-Pinene 4.2–4.4 13–32 n.m.c 58 n.m.
Limonene 4.4–4.5 12–13 n.m. 58 n.m.
b-Myrcene 4.3–4.9 <10 0.05 52 48
a-Humulene 7.0 N/A 0.20 64 36
Extractability values represent overall means across all treatments in which the analyte was detected.
a
Partition coefficient data from the literature [38–40].
b
Solubility data from the literature [37,41,42].
c
n.m.¼not measured.

Figure 2. Extraction into beer (white circles) and retention in spent hops (black circles) of sesquiterpene hydrocarbons (a-humulene, b-farnesene, b-caryophyllene),
monoterpene hydrocarbons (b-myrcene, a-, b-pinene, limonene, extraction only measured in b-myrcene), terpene esters (geranyl acetate), and terpene alcohols
(linalool, geraniol, nerol) as a function of log (KOW).

quantitatively abundant constituent in hop oil. The mean b-myrcene is one of the most potent odorants in hop cones,
retention rates (52–58%) of the monoterpene hydrocarbons its contribution to hoppy aroma in beer is less clear, as it is
(b-myrcene, a- and b-pinene, and limonene, log(Kow) ¼ nearly completely lost during wort-boiling, and is only
4.3–4.5) covered a similar range to the sesquiterpenes, but found above its aroma threshold in dry-hopped beer.[15,45]
was in general slightly lower. The terpene ester geranyl acet- Linalool, on the other hand, is often identified as a key
ate (log(Kow) ¼ 4.04) had a relatively higher mean extrac- aroma compound in both hops and beer.[13,45–47] The
tion rate (9%) and lower mean retention rate (36%), in line b-myrcene and linalool content of dry-hopped beer as well
with its slightly lower partition coefficient. The terpene alco- as their extraction and retention rates were plotted against
hols (linalool, geraniol, and nerol, log(Kow) ¼ 3.47–3.56) their initial concentration in the hop material, in order to
showed by far the highest mean extraction rates (43–51%) take a closer look at their behavior during dry-hopping
and the lowest mean retention rates (16–42%) owing to the (Figure 3). In both cases, statistically significant positive lin-
more polar character of the alcohol moiety. Generally speak- ear correlations were found between the dose of these com-
ing, with increasing Kow, an increase in retention rate with pounds and their concentration in beer (P < 0.0001 and
an accompanying decrease in extraction rate was observed. P ¼ 0.0006 for b-myrcene and linalool, respectively). In spite
It is also worth noting there were relatively high rates of the considerably smaller extraction rates seen for b-myr-
(>30%) of unaccounted for losses across all compound cene, both compounds showed a statistically significant
groups. It is possible that these losses could have occurred negative correlation between extraction rate and dosage
during filtration, although, to the best of our knowledge, (P < 0.0001 and P ¼ 0.0009 for b-myrcene and linalool,
this effect has not been demonstrated. respectively). This finding is important evidence that higher
dry-hop addition rates may lead to slight increases in aroma
potential, but come at the cost of diminishing returns in
Transfer, extraction, and retention of myrcene
terms of both hop volatile extraction, and hop aroma inten-
and linalool
sity; this phenomena has been documented by Lafontaine
The compound b-myrcene, an acyclic monoterpene hydro- and Shellhammer[9] and other researchers.[18]
carbon, is typically the most abundant constituent in hop In terms of retention rates, a significant correlation was
oil, making up 30–60% of the total oil content.[44] Although only seen for linalool (P < 0.0001), which showed a positive
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS 9

Figure 3. Concentration in beer (gray circles), extraction into beer (white circles), and retention in spent hops (black circles) for b-myrcene (left) and linalool (right)
as a function of effective dry-hopping dosage.

Figure 4. The b-myrcene (red) concentration as a function of linalool (blue) concentration in raw (black circles) and spent (white circles) hops on a moisture free
basis (db).

correlation between dosage and retention, indicating that than the raw hops, at 271 compared with 107, respectively.
higher addition rates lead to decreased efficiency. As a com- This indicates a preferential enrichment of b-myrcene in spent
parison of the relative extraction/retention rates of these two hops relative to linalool, which can be attributed to the differ-
compounds, b-myrcene content was plotted against linalool ence in their polarity and partitioning behavior. This simple
content of hops pre-dry-hopping and post-dry-hopping on an comparison draws attention to the likelihood that spent hops
“as is” dry basis (Figure 4). In this data set, the amounts of have more hydrocarbon character than their raw counterparts,
these two compounds were positively correlated and depended which could lead to differences in aroma quality upon their
on the quantity of total oil in the hop material. A linear regres- addition to beer. Additionally, this comparison points to the
sion was fit for the two sets of data, with intercepts forced likelihood that spent dry-hops are not a wholesale replacement
through the origin in order to compare the slopes and by for first-use hop material, and likely need to be used in a dif-
proxy the b-myrcene/linalool ratios (R2 ¼ 0.89 and 0.46 in ferent manner or during a different stage of the brewing pro-
raw and spent hops, respectively). The slope of the line of best cess in order to be utilized effectively from a flavor/aroma
fit for the spent hops was considerably higher in magnitude perspective.
10 D. G. HAUSER ET AL.

Conclusions and implications Literature cited


The results of this investigation provide proof that dry-hop- [1] Mitter, W.; Cocuzza, S. Revival of a Process (Dry Hopping-
ping is a relatively inefficient process, and that spent dry- Basics and Techniques). Brew. Bev. Ind. Int. 2013, 3, 28–30.
[2] Ting, P. L.; Ryder, D. S. The Bitter, Twisted Truth of the Hop:
hops waste on both a pilot and industrial scale contain a 50 Years of Hop Chemistry. J. Am. Soc. Brew. Chem. 2017, 75,
considerable quantity of both volatile and non-volatile hop 161–180. DOI: 10.1094/ASBCJ-2017-3638-01.
constituents. This opens the possibility of reclaiming both [3] Biendl, M.; Engelhard, B.; Forster, A.; Gahr, A.; Lutz, A.;
the bittering and aroma potential of what would otherwise Mitter, W.; Schmidt, R.; Schonberger, C. Hops: Their
Cultivation, Composition and Usage. Fachverlag Hans Carl:
be a waste material of the production process. The beer N€urnberg, 2014.
industry is an intensive user of energy and water, has a sig- [4] George, A. 2017 Statistical Report Presented by the Hop
nificant carbon footprint, and generates an appreciable Growers of America; 2017.
quantity of both solid and liquid waste.[23] Recent research [5] Malowicki, M. G.; Shellhammer, T. H. Factors Affecting Hop
Bitter Acid Isomerization Kinetics in a Model Wort Boiling
has demonstrated that in spite of the potential for unsus- System. J. Am. Soc. Brew. Chem. 2006, 64, 29–32. DOI:
tainable practices in brewing, many regional craft brewers 10.1094/ASBCJ-64-0029.
are succeeding in assessing the potential impacts of their [6] EBC Technology and Engineering Forum, Hops and Hop
production processes and implementing changes towards Products: (Manual of Good Practice). Getranke-Fachverlag Hans
Carl: N€urnberg, 1997.
better environmental, economic, and social sustainability.[48]
[7] Wolfe, P. H. A study of factors affecting the extraction of flavor
In addition, consumer research has shown that for low risk when dry hopping beer. Thesis - Master of Science. Oregon
products, consumers are often willing to pay more for those State University, 2012.
that contain recycled materials,[49] and that beer drinkers [8] Parkin, E.; Shellhammer, T. Toward Understanding the
Bitterness of Dry-Hopped Beer. J. Am. Soc. Brew. Chem. 2017,
are willing to pay more for more sustainably brewed
75, 363–368. DOI: 10.1094/ASBCJ-2017-4311-01.
beer.[24] Therefore, the recycling of spent dry-hops for either [9] Lafontaine, S. R.; Shellhammer, T. H. Impact of Static Dry-
bittering, late-, or dry-hops additions could provide an Hopping Rate on the Sensory and Analytical Profiles of Beer. J.
opportunity for brewers to simultaneously improve the eco- Inst. Brew. 2018, 124, 434–442. DOI: 10.1002/jib.517.
nomic/environmental sustainability of production and differ- [10] Briggs, D. E.; Boulton, C. A.; Brookes, P. A.; Stevens, R., Eds.
The Chemistry of Hop Constituents - Chapter 8 in Brewing
entiate themselves by offering a novel product. Science and Practice, Woodhead Publishing: Sawston, 2004;
It is important to note that the results of this study repre- pp 255–305.
sent distinct dry-hopping regimes, and the reader is cau- [11] Jaskula, B.; Kafarski, P.; Aerts, G.; De Cooman, L. A Kinetic
tioned in extending the findings to the wide range of dry- Study on the Isomerization of Hop Alpha-Acids. J. Agric. Food
Chem. 2008, 56, 6408–6415. DOI: 10.1021/jf8004965.
hopping methods used in industry. Further work also needs [12] De Keukeleire, D. Fundamentals of Beer and Hop Chemistry.
to be carried out in order to assess how well spent dry-hops Quım. Nova 2000, 23, 108–112. DOI: 10.1590/S0100-4042200
are utilized in kettle additions, and whether the use of spent 0000100019.
dry-hops for bittering (alone or in combination with raw [13] Schonberger, C.; Kostelecky, T. 125th Anniversary Review: The
Role of Hops in Brewing. J. Inst. Brew. 2011, 117, 259–267.
hops) produces an acceptable beer from an organoleptic per-
DOI: 10.1002/j.2050-0416.2011.tb00471.x.
spective. The same approach must be taken with the poten- [14] Roberts, M. T.; Dufour, J.-P.; Lewis, A. C. Application of
tial re-use of spent material for late- or dry-hopping, given Comprehensive Multidimensional Gas Chromatography
the notable enrichment in hydrocarbon character relative to Combined with Time-of-Flight Mass Spectrometry (GC  GC-
TOFMS) for High Resolution Analysis of Hop Essential Oil. J.
higher impact oxygenated compounds.
Sep. Sci. 2004, 27, 473–478. DOI: 10.1002/jssc.200301669.
[15] Rettberg, N.; Biendl, M.; Garbe, L.-A. Hop Aroma and Hoppy
Beer Flavor: Chemical Backgrounds and Analytical Tools—A
Acknowledgments Review. J. Am. Soc. Brew. Chem. 2018, 76, 1–20. DOI: 10.1080/
03610470.2017.1402574.
The authors would like to thank Hopsteiner and Virgil Gamache [16] Kishimoto, T.; Wanikawa, A.; Kono, K.; Shibata, K.
Farms for providing hop material, as well as Bridgeport Brewing for Comparison of the Odor-Active Compounds in Unhopped Beer
supplying the un-hopped base beer. The authors would also like to and Beers Hopped with Different Hop Varieties. J. Agric. Food
acknowledge the OSU hops laboratory technicians Lorena Ramos and Chem. 2006, 54, 8855–8861. DOI: 10.1021/jf061342c.
Nick Spatafora for providing support during the pilot-scale dry-hop- [17] Dresel, M.; Praet, T.; Opstaele, F.; Van Holle, A.; Naudts, D.;
ping and chemical analysis. De Keukeleire, D.; De Cooman, L.; Aerts, G. Comparison of the
Analytical Profiles of Volatiles in Single-Hopped Worts and
Beers as a Function of the Hop Variety. Brew. Sci. 2015, 68,
8–28.
Funding [18] Hanke, S.; Herrmann, M.; R€ uckerl, J.; Sch€
onberger, C.; Back,
This research was supported by the Hop Research Council and a Barth W. Hop Volatile Compounds (Part II): Transfer Rates of Hop
Haas Grant. Compounds from Hop Pellets to Wort and Beer. Brew. Sci.
2008, 61, 140–147.
[19] Forster, A.; Gahr, A. On the Fate of Certain Hop Substances
ORCID during Dry Hopping. Brew. Sci. 2013, 66, 93–103.
[20] Dieckmann, R. H.; Palamand, S. R. Autoxidation of Some
Dean G. Hauser http://orcid.org/0000-0001-9227-1139 Constituents of Hops. I. Monoterpene Hydrocarbon, Myrcene.
Scott R. Lafontaine http://orcid.org/0000-0003-4098-7711 J. Agric. Food Chem. 1974, 22, 498–503. DOI: 10.1021/
Thomas H. Shellhammer http://orcid.org/0000-0002-4055-2366 jf60193a033.
JOURNAL OF THE AMERICAN SOCIETY OF BREWING CHEMISTS 11

[21] Cocuzza, S.; Mitter, W. Dry Hopping – a Study of Various [36] Yaws, C. L. Octanol-Water Partition Coefficient for Organic
Parameters. Brew. Bev. Ind. Int. 2013, 4, 70–74. Compounds - Chapter 16. In Yaws’ Handbook of Properties for
[22] Haslbeck, K.; Minkenberg, D.; Coelhan, M. Investigations into Aqueous Systems. Knovel: Norwich, NY, 2012.
the Transfer Rate of Volatile Compounds in Dry Hopping [37] Martins, M. A. R.; Silva, L. P.; Ferreira, O.; Schr€ oder, B.;
Using an Octanol-Water Partition Coefficient Model. J. Am. Coutinho, J. A. P.; Pinho, S. P. Terpenes Solubility in Water
Soc. Brew. Chem. 2018, 76, 169–177. DOI: 10.1080/03610470. and Their Environmental Distribution. J. Mol. Liq. 2017, 241,
2018.1483701. 996–1002. DOI: 10.1016/j.molliq.2017.06.099.
[23] Olajire, A. A. The Brewing Industry and Environmental [38] Griffin, S.; Wyllie, S. G.; Markham, J. Determination of
Challenges. J. Clean. Prod. 2012, 1–21. DOI: 10.1016/ Octanol–Water Partition Coefficient for Terpenoids Using
j.jclepro.2012.03.003. Reversed-Phase High-Performance Liquid Chromatography. J.
[24] Carley, S.; Yahng, L. Willingness-to-Pay for Sustainable Beer. Chromatogr. A 1999, 864, 221–228. DOI: 10.1016/S0021-
PLoS One. 2018, 13, e0204917DOI: 10.1371/journal.pone. 9673(99)01009-2.
0204917. [39] Kishimoto, T.; Wanikawa, A.; Kagami, N.; Kawatsura, K.
[25] Lafontaine, S. R.; Varnum, S.; Roland, A.; Delpech, S.; Dagan, Analysis of Hop-Derived Terpenoids in Beer and Evaluation of
L.; Vollmer, D.; Kishimoto, T.; Shellhammer, T. H. Impact of Their Behavior Using the Stir Bar-Sorptive Extraction Method
Harvest Maturity on the Aroma Characteristics and Chemistry with GC-MS. J. Agric. Food Chem. 2005, 53, 4701–4707. DOI:
of Cascade Hops Used for Dry-Hopping. Food Chem. 2019, 10.1021/jf050072f.
278, 228–239. DOI: 10.1016/j.foodchem.2018.10.148. [40] Copolovici, L. O.; Niinemets, U. Temperature Dependencies of
[26] Lafontaine, S. R.; Hauser, D. G.; Foster, R.; Donaldson, J.; Henry’s Law Constants and Octanol/Water Partition
Gamache, D.; Shellhammer, T. H. Impact of Kiln Temperatures Coefficients for Key Plant Volatile Monoterpenoids.
on the Aroma and Enzymatic Potential of Hops during Dry- Chemosphere. 2005, 61, 1390–1400. DOI: 10.1016/
Hopping. Recent Advances in Hop Science, European Brewing j.chemosphere.2005.05.003.
Congress: N€ urnberg, 2018. [41] Weidenhamer, J. D.; Macias, F. A.; Fischer, N. H.; Williamson,
[27] Vollmer, D. M.; Shellhammer, T. H. Dry Hopping on a Small
G. B. Just How Insoluble Are Monoterpenes? J. Chem. Ecol.
Scale: Considerations for Achieving Reproducibility. Tech. Q.
1993, 19, 1799–1807. DOI: 10.1007/BF00982309.
Master Brew. Ass. Am. 2016, 53, 140–144.
[42] Tamura, K.; Li, H. Mutual Solubilities of Terpene in Methanol
[28] American Society of Brewing Chemists, Hops – 4a, Moisture by
and Water and Their Multicomponent Liquid  Liquid
Routine Air Oven Method, -14, a-Acids and b-Acids in Hops
Equilibria. J. Chem. Eng. Data 2005, 50, 2013–2018. DOI:
and Hop Extracts by HPLC (International Method), – 13, Total
10.1021/je0502391.
Essential Oils in Hops and Hop Pellets by Steam Distillation, –
[43] Praet, T.; Van Opstaele, F.; Baert, J.; Aerts, G.; De Cooman, L.
17, Hop Essential Oils by Capillary Gas Chromatography–Flame
Comprehensive Characterisation of the Hop-Derived
Ionization Detection; Beer 23E, Iso-a-Acids in Beer by HPLC, 8
ed.; American Society of Brewing Chemists: St Paul, MN, 1992. Sesquiterpenoid Fingerprint of American Kettle Hopped Lager
[29] Kaneko, T.; Ito, K.; Izumi, T.; Matsuyama, A. Determination of Beers. Brew. Sci. 2014, 67, 183–194.
Alcohol and Real Extract by Anton Paar Alcolyzer Y1 - 2006/ [44] Almaguer, C.; Sch€ onberger, C.; Gastl, M.; Arendt, E. K.; Becker,
09/01. J. Am. Soc. Brew. Chem. 2006, 64, 252–254. T. Humulus lupulus-A Story That Begs to Be Told. A Review. J.
[30] Lafontaine, S. R.; Shellhammer, T. H. Sensory Directed Mixture Inst. Brew. 2014, 120, 289–314.
Study of Beers Dry-Hopped with Cascade, Centennial, and [45] Steinhaus, M.; Schieberle, P. Comparison of the Most Odor-
Chinook. J. Am. Soc. Brew. Chem. 2018, 76, 199–208. DOI: Active Compounds in Fresh and Dried Hop Cones (Humulus
10.1080/03610470.2018.1487747. lupulus L. Variety Spalter Select) Based on GC  Olfactometry
[31] Sharp, D. C.; Townsend, M. S.; Qian, Y.; Shellhammer, T. H. and Odor Dilution Techniques. J. Agric. Food Chem. 2000, 48,
Effect of Harvest Maturity on the Chemical Composition of 1776–1783. DOI: 10.1021/jf990514l.
Cascade and Willamette Hops. J. Am. Soc. Brew. Chem. 2014, [46] Hanke, S. Linalool—A Key Contributor to Hop Aroma. MBAA
72, 231–238. DOI: 10.1094/ASBCJ-2014-1002-01. Global Emerging Issues 2009. https://www.mbaa.com/brewre-
[32] Cosbie, A. J. C. The Effect of Temperature of Drying on the sources/Documents/Linalool.pdf
Value of Hops. J. Inst. Brew. 1937, 43, 9–16. DOI: 10.1002/ [47] Vollmer, D. M.; Lafontaine, S. R.; Shellhammer, T. H. Aroma
j.2050-0416.1937.tb05708.x. Extract Dilution Analysis of Beers Dry-Hopped with Cascade,
[33] Hops Advisory Committee. Rosemaund Hop Drying Chinook, and Centennial. J. Am. Soc. Brew. Chem. 2018, 76,
Experiment with 1959 and 1960 Crop Hops. J. Inst. Brew. 1962, 190–198. DOI: 10.1080/03610470.2018.1487746.
68, 395–397. DOI: 10.1002/j.2050-0416.1962.tb01881.x. [48] Hoalst-Pullen, N.; Patterson, M. W.; Mattord, R. A.; Vest,
[34] Maye, J. P.; Smith, R. Hidden Secrets of the New England IPA. M. D. Sustainability Trends in the Regional Craft Beer
Tech. Q. Master Brew. Ass. Am. 2018, 55, 88–92. Industry. In The Geography of Beer, Hoalst-Pullen, N.;
[35] Vollmer, D. M.; Shellhammer, T. H. Influence of Hop Oil Patterson, M. W. Eds. Springer: Dordrecht, 2014; pp 109–116.
Content and Composition on Hop Aroma Intensity in Dry- [49] Hamzaoui Essoussi, L.; Linton, J. D. New or Recycled Products:
Hopped Beer. J. Am. Soc. Brew. Chem. 2016, 74, 242–249. DOI: How Much Are Consumers Willing to Pay? J. Consum. Market.
10.1094/ASBCJ-2016-4123-01. 2010, 27, 458–468. DOI: 10.1108/07363761011063358.

View publication stats

You might also like