Processes 10 00071

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

processes

Review
The Potential of Spectroscopic Techniques in Coffee
Analysis—A Review
Leah Munyendo 1,2, *, Daniel Njoroge 2 and Bernd Hitzmann 1

1 Department of Process Analytics and Cereal Science, University of Hohenheim, Garbenstr. 23,
70599 Stuttgart, Germany; bernd.hitzmann@uni-hohenheim.de
2 Institute of Food Bioresources Technology, Dedan Kimathi University of Technology, Nyeri 10143, Kenya;
daniel.njoroge@dkut.ac.ke
* Correspondence: leah.munyendo@uni-hohenheim.de; Tel.: +49-1573-2373354

Abstract: This review provides an overview of recent studies on the potential of spectroscopy
techniques (mid-infrared, near infrared, Raman, and fluorescence spectroscopy) used in coffee
analysis. It specifically covers their applications in coffee roasting supervision, adulterants and
defective beans detection, prediction of specialty coffee quality and coffees’ sensory attributes,
discrimination of coffee based on variety, species, and geographical origin, and prediction of coffees
chemical composition. These are important aspects that significantly affect the overall quality of
coffee and consequently its market price and finally quality of the brew. From the reviewed literature,
spectroscopic methods could be used to evaluate coffee for different parameters along the production
process as evidenced by reported robust prediction models. Nevertheless, some techniques have
received little attention including Raman and fluorescence spectroscopy, which should be further
studied considering their great potential in providing important information. There is more focus
on the use of near infrared spectroscopy; however, few multivariate analysis techniques have been
explored. With the growing demand for fast, robust, and accurate analytical methods for coffee
 quality assessment and its authentication, there are other areas to be studied and the field of coffee

spectroscopy provides a vast opportunity for scientific investigation.
Citation: Munyendo, L.; Njoroge, D.;
Hitzmann, B. The Potential of
Keywords: spectroscopy; coffee; chemometrics; prediction; adulterants; defective beans; discrimination;
Spectroscopic Techniques in Coffee
roasting
Analysis—A Review. Processes 2022,
10, 71. https://doi.org/10.3390/
pr10010071

Academic Editor: Dariusz Dziki 1. Introduction


Received: 30 November 2021 Coffee is one of the most widely traded commodities worldwide and is produced
Accepted: 27 December 2021 in over 50 countries. According to ICO [1], world green coffee production reached
Published: 30 December 2021 10.52 million tonnes in 2020, a 6.3% increase compared to 2019. Generally, production
Publisher’s Note: MDPI stays neutral
has been increasing year by year, clearly indicating an upward trend in coffee consumption.
with regard to jurisdictional claims in
This can be attributed to coffees’ desirable sensory properties, the stimulant effects of
published maps and institutional affil- caffeine, as well as its benefits on human health. More than 100 species in the genus Coffea
iations. have been identified; however, only two species, namely, Coffea canephora and Coffea arabica,
commonly known as robusta and arabica, respectively, are traded commercially. Arabica
coffee accounts for more than 60% of the world’s coffee production with a higher commer-
cial value than robusta coffee due to its superior aroma characteristics and lower caffeine
Copyright: © 2021 by the authors. content [2].
Licensee MDPI, Basel, Switzerland. Assessment of coffee along its supply chain, i.e., from farm to cup, is crucial in
This article is an open access article ensuring consistent supply of quality products to consumers. The overall quality of a
distributed under the terms and cup of coffee is directly linked to careful control of a number of parameters including the
conditions of the Creative Commons quality of cherries, raw materials (species, origin, presence of adulterants and defective
Attribution (CC BY) license (https://
beans, type of processing method (wet or dry), chemical composition etc.), roasting process,
creativecommons.org/licenses/by/
storage conditions, and preparation parameters and methods. In the past years, different
4.0/).

Processes 2022, 10, 71. https://doi.org/10.3390/pr10010071 https://www.mdpi.com/journal/processes


Processes 2022, 10, 71 2 of 25

analytical techniques have been developed and used to authenticate and assess the quality
of coffee. They may involve target analyses, in which chemical profiling is done and the
identified compounds are assessed against a set limit established by law or stated on the
label. Non-target analyses, in contrast, provide a fingerprint of the whole sample that can
then be utilized to detect possible adulteration. Physicochemical and chromatographic
analytical methods are the predominant methods in the evaluation of coffee species, origin,
and chemical composition [3,4]. Although very objective, these methods usually require
experienced personnel, and are labor intensive, expensive, and time consuming. Visual
coffee inspection is also a common practice that is done to check for the presence of defective
beans, adulterants, degree of roast as well as to distinguish coffee species based on the
colour, shape, and size of the bean. This is a subjective method that depends mainly on the
experience and skills of the inspector and hence could be subject to erroneous conclusions.
The global coffee market is increasingly stressing the importance of the quality and
chemical characteristics of coffee, which calls for appropriate analytical methods. Spec-
troscopy is a promising technique in this industry because it is rapid, non-destructive,
requires no/minimal sample preparation, and integrates easily into other processes. Infra-
red spectral techniques have gained considerable interest in evaluation of coffee with
respect to its chemical composition, species, geographical origin, presence of adulterants
and defective beans, roast degree, sensory attributes, etc. [5–10]. The potential of Raman
and fluorescence spectroscopy in coffee analysis have also been investigated [11–14].
Potential application of spectroscopy techniques in the food industry has gained
interest over the last few years. Therefore, the aim of this review is to provide a summary
of the recent studies on applications of near infrared, fluorescence, mid infrared, and
Raman spectral techniques in coffee assessment. It is hoped that this will provide an
overview to interested parties on what can be achieved using spectroscopy techniques in
coffee analysis. Additionally, aspects that still require clarification or further investigations
are identified. However, this review does not address all the aspects related to coffee
assessment. Important topics such as prediction of coffee processed using different methods,
differentiation of organic from inorganic coffee, monitoring of coffee quality during storage,
among others, were not considered. Limited information is available, and therefore these
are areas that could be considered in the future given their significant impact on the
quality of the beverage. In this review, first a brief introduction to the basic concepts of the
aforementioned spectroscopic methods is provided followed by discussion of the latest
research carried out to develop new approaches based on these methods.

2. Overview of Spectroscopic Techniques


Spectroscopic techniques have gained interest in the food industry over the past
years due to their potential in assessing the chemical composition and quality of foods.
These techniques involve interaction between matter and electromagnetic radiation, which
provides useful information about food samples and can be used for both qualitative
and quantitative analysis. These methods present some advantages over conventional
analytical methods given their rapidity, provision of on-line analysis, minimal or no sample
preparation, and the fact that they are non-destructive [15]. Fluorescence spectroscopy
techniques utilizes a beam of light, which excites the electrons in molecules of certain
compounds commonly known as fluorophores, causing them to emit light. At a specific
wavelength, fluorophores will absorb energy in the form of light and later liberate this
energy in the form of light emission at a longer wavelength.
The processes that occur between light absorption and emission are clearly explained
in Figure 1 for fluorescence, NIR, and Raman spectroscopy. In fluorescence (Figure 1a)
visible light is involved, where the electronic state of a molecule is changed. In the figure,
S0 and S1 denote the singlet ground and first excited electronic state, respectively. The
fluorescence process starts with the excitation where a molecule absorbs light and gets
excited from ground state (S0 ) to a higher vibrational level in S1 . This is followed by
rapid relaxation of molecules from an upper vibrational level to the lowest level in the
orescence process starts with the excitation where a molecule absorbs light
cited from ground state (S0) to a higher vibrational level in S1. This is follow
relaxation of molecules from an upper vibrational level to the lowest level in
Processes 2022, 10, 71 3 of 25
tronic excited state S1, without any radiation. This process is known as interna
and normally happens within 10−12 s or less. Finally, emission occurs usuall
excitation, when the molecule goes back to any vibrational state of its electr
first electronic excited state S1 , without any radiation. This process is known as internal
state, S0. In modern food analysis, the fluorescence spectroscopy method has
conversion and normally happens within 10− 12 s or less. Finally, emission occurs usually
sively used in the evaluation of food components, adulterants, contaminan
10− 8 s after excitation, when the molecule goes back to any vibrational state of its electronic
ground state, tives, thanks to
S0 . In modern fooditsanalysis,
high specificity and sensitivity
the fluorescence [16,17].
spectroscopy method has been
extensively used in the evaluation of food components, adulterants, contaminants, and
additives, thanks to its high specificity and sensitivity [16,17].

a) b) c)
Electronic excited state Mostly non symmetric, Mostly symmetric, non
polar molecules polar molecules
Virtual level
Electronic ground state

S1

S0 Difference

Fluorescence NIR Raman


4000 -12,000 cm-1 200 -4000 cm-1
270 - 600 nm 800 -2500 nm
Electronic transition Mostly Mostly ground
overtones oscillation

Figure 1. Simplified potential curves to illustrate different spectroscopic methods discussed here; just
vibrational states are presented: (a) fluorescence, (b) NIR, (c) Raman. The red arrow indicates the
energy involved in the corresponding radiation.
Figure 1. Simplified potential curves to illustrate different spectroscopic methods d
Infrared just
(IR)vibrational
spectroscopy states is among the most
are presented: (a) common
fluorescence,vibrational
(b) NIR,spectroscopic
(c) Raman. The red ar
methods usedtheinenergythe food industry.
involved in theItcorresponding
is usually divided into three regions of wave-
radiation.
length: far infrared (15 µm–1 mm or 400–10 cm− 1 ), mid infrared (MIR: 4000–400 cm− 1
or 2500–25,000 nm),Infrared
and near(IR) infrared (NIR:13,333–4000
spectroscopy is among cm−the1 or 800–2500 nm) [18]. IR
most common vibrational s
spectroscopy methods
is based on thein
used principle
the foodthat every chemical
industry. bonddivided
It is usually vibratesintowiththree
specific
regions of
frequency, which corresponds to certain amount of energy.−1 These vibrations induce ab-
far infrared (15 µm–1 mm or 400–10 cm ), mid infrared (MIR: 4000–400 c
sorption bands, useful in the qualitative and quantitative analyses of−1many molecules
25,000 nm), and near infrared (NIR:13,333–4000 cm or 800–2500 nm) [18].
in samples. Vibrations of certain functional groups, for instance C=O, NH2 , –OH, –CH3 ,
copy produce
C6 H5 – etc., always is basedbands
on theinprinciple
the infrared that every chemical
spectrum bond vibrates
within frequency with specif
ranges that
are well-definedwhich corresponds
irrespective to certain
of the molecule amount
having of energy.
the functional These
group. vibrations induc
Identification
and designation bands,
of theseuseful
bandsin to the qualitative
particular and quantitative
chemical groups provide analyses of many molecule
specific information
about the product being investigated
Vibrations [17,19].
of certain functional groups, for instance C=O, NH2, –OH, –CH
Broad bands rising from combinations
always produce bands in the andinfrared
overtones of the fundamental
spectrum vibrationsranges t
within frequency
relating to O–H, N–H,irrespective
defined and C–H chemical of the bonds
molecule characterize
having the spectra in the NIR
functional region
group. Identificat
(Figure 1b). Samples with different chemical bonds, i.e., chemical composition, will record
ignation of these bands to particular chemical groups provide specific inform
different spectra, which are unique to them acting as a ‘fingerprint’. NIR spectroscopy has
the product being investigated [17,19].
been widely used for analysing the quality and chemical composition of foods including
Broad
meat, fish, milk, cereals, bands
fruits, rising
cereals, from
wine, beer,combinations
eggs, and theirand overtones
related productsof the fundament
[17,19,20].
relating
Despite its great potential to in
O–H,
foodN–H, and
analysis, theC–H chemicalofbonds
interpretation characterize
the spectra spectra in the
in NIR analysis
is challenging(Figure
due to its 1b). Samples with
broadband nature,different chemical
which consists of bonds,
overlappingi.e., chemical
overtone composition
and
combination bands.
different Specific compounds
spectra, which are are unique
not as well determined
to them actingasasthey are in the MIR
a ‘fingerprint’. NIR spec
region. This hasbeenhowever,
widelybeen used addressed by datathe
for analysing processing
quality using chemometric
and chemical methods of foo
composition
to relate spectral
meat,information
fish, milk, to properties
cereals, of the samples
fruits, cereals,[21].
wine, beer, eggs, and their relat
[17,19,20]. Despite its great potential in food analysis, the interpretation of t
Processes 2022, 10, 71 4 of 25

Mid-infrared spectroscopy provides information on the related rotational-vibrational


structure as well as fundamental vibrations of the chemical bonds in the samples. Through
technological advances, original MIR spectrophotometers, which measured the amount
of absorbed energy based upon the fact that the infrared light frequency varies when
passing through a monochromator, have been replaced by Fourier transform infrared (FTIR)
spectrophotometers. FTIR equipment uses interferometers instead of monochromators,
yielding advantages such as high sensitivity and precision, fast analysis, and a high signal-
to-noise ratio. Consequently, the term “FTIR spectroscopy” and “MIR spectroscopy” are
sometimes interchangeably used [22], although FTIR is also applied for NIR. The mid-
infrared spectrum is comprised of four regions: the triple bond, the X–H stretching, the
fingerprint, and the double bond regions. Fundamental vibrations from O–H, C–H, and
N–H stretching characterize the X–H stretching region. The triple bond stretching region
is generally due to C≡N and C≡C bonds vibrations, whereas in the double bond region,
absorption bands characterized by C=N, C=C, and C=O occur. Bending and skeletal
vibrations characterize the fingerprint region [17]. Utilization of MIR spectroscopy has
been used as an effective analytical method in the food industry [10,23]. The far infrared
(far-IR) spectral region (10–400 cm− 1 ) is the range of wavenumbers characterized by large-
amplitude anharmonic vibrations. The technique based on the far-IR has shown to be more
suitable in providing information on dynamics of proteins and analyzing structures that are
highly ordered, for instance, the formation of fibrillar, due to its sensitivity to vibrational
modes resulting from hydrogen bonds and peptide skeletons [24].
In Raman spectroscopy (Figure 1c), a virtual excited energy state is involved. Compar-
ing Raman spectroscopy to IR spectroscopy, the information provided by Raman is due to
inelastic light scattering, as opposed to IR, which relies on elastic scattering. In contrast to
IR, Raman spectra are obtained from symmetric and non-polar molecules. Raman scattering
results from inelastic scattering of the incident photons through which energy is received
from or transferred to the molecule owing to the changes in the vibrational or rotational
modes of sample molecules. This causes a change in the energy, and thus the frequency of
the scattered light [25].
A Raman spectrum is a plot of scattering intensity vs. wavenumber or wavelength
consisting of bands, which corresponds to a Raman shift that occurs because of the dif-
ference in excitation and emission light energy (see Figure 1c). So-called Stokes–Raman
lines are due to smaller emission energy than excitation, and for Anti-Stokes–Raman lines
this behavior is reversed, but with much smaller intensity. The obtained Raman bands
characterize specific functional groups and/or chemical bonds (C–X (X = Cl, I, F, or Br),
C–S, C–NO2 , S–S, C=N, C=C) of materials and by using them, it is possible to obtain
fingerprints of materials with Raman spectroscopy. As the bands’ intensity is directly
proportional to the analyte concentration of the materials, quantitative analyses using
Raman spectroscopy is possible [26]. Analyses of safety, quality, and composition of foods
are among the applications of Raman spectroscopy in the food industry [11,27,28].
In summary, the potential of spectroscopic methods in the food industry is enormous
and broad. With regard to coffee, there is much literature on the utilization of these
techniques in determining coffees authenticity, quality, geographical origin, species, and
composition and detecting misrepresentation and adulteration. Table 1 shows reported
absorption bands of coffees chemical components in the IR and Raman region.
Processes 2022, 10, 71 5 of 25

Table 1. Reported absorption bands of coffees chemical components in the IR and Raman region.

Wavenumber (cm−1 ) or
Vibrational Modes Associated Compounds References
Wavelength (nm)
NIR
4750 and 6900 and 5200 cm−1 - Proteins and water, respectively
[29]
4200 to 4400 cm−1 - Carbohydrates
1st overtone of
1450 nm Water
O–H stretching
2nd overtone of Carbohydrates, quinic acid,
1100 to 1250 nm
C–H stretching and lipids [30]
1300–1350 nm - Caffeine
CGA, carbohydrates, and
1550 nm -
amino acids
5000–5200 cm−1 and
- Water
6800–7200 cm−1
[31]
CGA, carbohydrates, proteins,
4000–5000 cm−1 C-H bonds vibrations
and trigonelline
3rd overtone of the CH, Ferulic and
900–1000 nm
CH2 , and CH3 groups coumaric acids
[32]
1st overtone of the OH
1400–1500 nm CGA, water, and carbohydrates
functional group
1st overtone of O–H
stretching and O–H
6750–6950 cm−1 and
deformation + O–H Water
5100–5200 cm−1
stretching combination [33]
bands
Lipids and CGA, proteins, caffeine,
5680–5850 and 4760 cm−1 -
and carbohydrates, respectively
1st overtone of C=O CGA, proteins, lipids, caffeine,
1850–1950 nm [34]
bonds vibrations water, and carbohydrates
Overtone and combination
4000–12,000 cm−1 bands of the NH, C–H, SH Proteins (amino acids) and lipids [35]
and OH bonds
Caffeine, fatty acids, amino acids,
4000–5600 cm−1 C-H bonds vibrations
and lignin
Fatty acids, amino acids,
8200–8300 cm−1 C-H bonds vibrations [36]
and lignin
10,000, 6800, 4400 and O-H and C-H
Cellulose
4800 cm−1 bond vibrations
1410, 1742, 1904, and 2318 nm - Lipids
1410, 1728, 1904, 2306 and
- Caffeine [37]
2348 nm
1436, 1880, 2312, 2324 and
- CGA
2350 nm
5000–5200 and
- Water
6800–7200 cm−1
1st overtone and [38]
5325, 5140, 5075, 5410–5385 combination band of C=O Carbohydrates, proteins, and
and 4980–5000 cm−1 and O-H bonds, 2nd organic acids
overtone of C=O
Processes 2022, 10, 71 6 of 25

Table 1. Cont.

Wavenumber (cm−1 ) or
Vibrational Modes Associated Compounds References
Wavelength (nm)
1868 nm - 5-CQA
1741 and 2252 nm [39]
2383–2488, 1977–2266, - Caffeine
1638–1793 and 1297–1474 nm
C–H + CC combination
2306–2312 and 1218–1242 nm bands and 2nd overtone of Sucrose and other carbohydrates
C–H, respectively
1472–1478 nm 1st overtone of N–H Phenols and CGA
O–H 1st overtone and [40]
2128–2132 and 2190–2192 nm Proteins and CGA
combination bands of N–H
Combination bands of
1706–1714, 2436–2475 and
C–H + C–H and C–H + CC Lipids
2480–2488 nm
and C-H 1st overtone
O–H stretching 1st
overtone and combination
1440–1480 and 1930–1950 nm Water
band of O–H deformation + [41]
O–H stretching
1715–1760 and 2300–2350 nm - Lipids
1180–1262 nm - Caffeine and cellulose
1480–1882 nm and CGA, caffeine, lipids, proteins,
- [5]
1896–2180 nm and carbohydrates
Caffeine, carbohydrates,
2260–2498 nm -
and protein
C-H stretching
1208 nm Sucrose, lipids, and amino acids
2nd overtone
1800–2000, 1400, 2200 and 1st overtones of O–H
Water
2300 nm and C-H
1st overtone of aromatic [42]
Around 1500 nm Phenolic compounds (CGA)
structures C–H
1700 and 1800 nm C-H 1st overtone Lipids, caffeine, and carbohydrates
C-H + CH2
2300–2400 nm Lipids
combination bands
O-H deformation and
6896 and 5154 cm−1 stretching Water
[43]
combination bands
4100–4400 cm−1 - CGA and quinic acid
MIR/
FTIR
Stretching of CH bonds in
2840–2940 cm−1 Lipids and caffeine
CH2 and CH3 groups
C=O stretch of aliphatic [41]
1747 cm−1 Lipids
ester groups
900–1400 cm−1 - Carbohydrates
Processes 2022, 10, 71 7 of 25

Table 1. Cont.

Wavenumber (cm−1 ) or
Vibrational Modes Associated Compounds References
Wavelength (nm)
Stretching of C=O and Carbohydrates and
1740 and 1660 cm−1
C=C bonds lipids, respectively
1600, 1700, 1260, 1160 and
- CGA [44]
1060 cm−1
1330 and 600–1300 cm−1 - Trigonelline and Pyridine
1020 cm−1 - Carbohydrates
Asymmetric and
symmetric stretching of
2922, 2852 and 1743 cm−1 Lipids
CH2 , and stretching of
C=O, respectively
[45]
Polysaccharides and other
1153 and 950–1130 cm−1 -
carbohydrates, respectively
C=O stretching
1643 cm−1 Caffeine
vibrations
650–900 cm−1 - Proteins and amino acids
Carbohydrates, CGA, proteins,
1000 and 1300 cm−1 - [46]
amino acids
1700–1750 and
C=O stretching Carbohydrates and lipids
1600–1680 cm−1
3470 cm−1 O-H stretching CGA and Water
Stretching of the C=C and Lipids and polysaccharides
3008 and 2855–2920 cm−1
C-H bond, respectively (lignin), respectively
Stretching and vibration of [47]
Quinic acids and fatty
1746 and 1704 cm−1 the ester group OC=O and
acids, respectively
carbonyl group C=O
Vibrations of the
1608 cm−1 Trigonelline and caffeine
C-N group
1745 cm−1 - Triglyceride
835, 911 and 1061 cm−1 - Polysaccharides
[48]
1268, 1291, 1322, 1337, 1374,
1396, 1398, 1406, 1418, and - Proteins and organic acids
1499 cm−1
1656 and 2800–3000 cm−1 - Caffeine
Lipids and caffeine or/and [49]
1743–1741 and 1543 cm−1 -
trigonelline
1744, 1654, and 1603 cm−1 - Lipid and caffeine
1285, 900–1200 and [50]
- CGA and carbohydrates
1400–1500 cm−1
804, 991, and 1020 cm−1 - Carbohydrates
Caffeine, triglycerides and [10]
1661, 1744 and 2922 cm−1 -
lipids, respectively
1600–1650 and
- Caffeine and CGA, respectively
1150–1450 cm−1
[8]
CGA, carbohydrates, and
600–1700 cm−1 -
trigonelline
Processes 2022, 10, 71 8 of 25

Table 1. Cont.

Wavenumber (cm−1 ) or
Vibrational Modes Associated Compounds References
Wavelength (nm)
CGA and pyruvic acid, pyridine,
3356 and 1067 cm−1 -
and quinic acid, respectively [51]
1000 and 1750 cm−1 - Caffeine and trigonelline
1744 and 1285 cm−1 - Esters and CGA, respectively
1600–1650 and 900–1200, Caffeine and [52]
-
1400–1500 cm−1 carbohydrates, respectively
Raman
Cafestol and
1500 and 1567 and 1478 cm−1 C=C stretch vibrations [53]
Kahweol, respectively
1600 cm −1 Phenyl ring stretch
1630 cm−1 C=C stretch vibrations
CH and COH [11]
1120 cm−1 Lipids, CGA, and proteins
bending vibration
Phenyl ring
1200 cm−1
bending vibration
Aromatic and
1604 and 1630 cm−1 Polyphenols, e.g., CGA
C=C stretching
Amide I band stretching [54]
1690 and 1656 cm−1 (structures of R-helix Proteins
and β-sheet)
1479 and 1567 cm−1 - Kahweol
C=C ethylenic stretch and
1630 and 1605 cm−1 Phenyl ring stretch CGA
vibrations, respectively
1000–1750 cm−1 - CGA [14]
CH2 asymmetric and
2934 and 2905 cm−1 Lipids
symmetric vibrations
2700–3050 cm−1 - Lipid bands
Cafestol and
1507 and 1485 and 1570 cm−1 - [55]
Kahweol, respectively
C=C and C=O
1606, 1637, 1657 and 1680 cm−1 CGA
bonds vibration
1567 and 1478 cm−1 - Kahweol
C-H bond symmetric and
2900 cm−1 Fatty acids [56]
asymmetric stretching
C=C deformation
1441, 1304 and 1265 cm−1 vibrations of CH, CH2 , and Fatty acids
CH3 bonds
1502 and 1442 cm−1 - Cafestol and lipids, respectively

3. Application of Spectroscopy Techniques in Coffee Analysis


3.1. Coffee Roasting and Monitoring
Roasting is a significant step in coffee processing where physical, chemical, and struc-
tural transformations occur. Moreover, there is production of melanoidins and thousands
of flavour compounds, all of which are accountable for the characteristic colour, aroma,
and flavour of the beverage [57]. Therefore, roasting is a step with a significant impact
Processes 2022, 10, 71 9 of 25

on the quality of the final product, thus warranting its control. Several studies on moni-
toring and controlling the coffee roasting process using spectroscopic methods have been
conducted (Table 2). Bertone et al. [58] studied the ability of Fourier transform (FT-NIR)
spectroscopy to determine roast degree (best on colour) and varietal composition of coffee
simultaneously. In the study, 130 blends of roasted-ground coffee containing both arabica
and robusta species at different levels were used. The partial least square regression (PLS)
algorithm was used, and validation was done with a completely independent external set.
The developed model had a root mean square error of prediction (RMSEP) of 1.28 Au for
colour and 4.34% (w/w) for arabica content in the blends, thus demonstrating a fast and
reliable assessment of important coffee characteristics.

Table 2. Summary of reported studies on potential applications of NIR, FTIR/MIR, fluorescence and
Raman spectroscopy in the coffee industry.

Spectroscopy Multivariate
Application Aim References
Technique Analysis
Coffee roasting Determine roast degree (based
NIR PLS [6,7,58,59]
and monitoring on color, acidity, cracks)
Online acidity monitoring
NIR PLS [38]
during coffee roasting
In-line coffee roasting
NIR PLS [31]
process monitoring
Real-time detection of faults
NIR PCA [60]
during coffee roasting process
Identify and quantify
Adulterants and adulterants (corn, barley, peels, PCA, PLS,
defective beans sticks, coffee husks, soy, ort, rice, NIR MCR-ALS and [32,47,61,62]
detection sticks, soil, and robusta) in SIMCA
roasted-ground coffee
Detect and quantify
adulterations of roasted-ground FTIR/MIR PCA and PLS [63]
coffee with corn
Quantify defective beans (black,
immature, dark sour and light FTIR/MIR and NIR PLS and Elastic net [41]
sour) in roasted coffees
Discriminate between defective
PCA, PLS,
(sour, immature, insect damaged,
NIR CEM-SVM and [64,65]
and black) and nondefective
CNN
green coffee beans
Discrimination of mature, sour,
AHC, LDA and
black, and immature green FTIR/MIR [49,66]
PCA
coffee beans
Evaluation of defects in
MIR/FTIR-PAS PLS-DA [51]
roasted-ground coffee
Prediction of specialty
coffee quality and Determine sensory attributes
NIR PLS, SNM, DCNN [40,67,68]
sensory attributes of coffee
of coffee
Prediction of specialty coffee
PLS, FFBPANN,
quality and its discrimination NIR [33,69,70]
CACHAS
from ordinary coffee
Prediction of coffee cup quality FTIR/MIR PCA, PLS-DA [45,71]
Quantitative evaluation of
sensory characteristics of
FTIR/MIR PLS, FA [8,44]
specialty coffees and
its discrimination
Discrimination of
Fluorescence PCA and SIMCA [12]
specialty coffee
Processes 2022, 10, 71 10 of 25

Table 2. Cont.

Spectroscopy Multivariate
Application Aim References
Technique Analysis
Discrimination of Potential function
Distinguish between robusta and
coffee based on class-modelling,
arabica coffee species and NIR [72–74]
species, variety, and PCA, SOM, SIMCA,
their varieties
geographical origin PLS-DA and SVM
PLS-DA, SNM,
Discriminate coffee of
NIR SIMCA, ANN [5,34,36,37,52,73]
different origin
(SOM)
ELM, BPNN,
Coffee variety identification FTIR/MIR RBFNN, RVM, [48]
SIMCA and SVM
Discriminate coffee of PLS, SVM, RBFs,
FTIR/MIR [50,52,73,75]
different origin MLP, SIMCA
Discrimination of coffee species PCA, MDA, QDA,
(arabica and robusta) Raman RDA, PLS–DA, [11,14,53–56]
and cultivars SIMCA, LDA
PARAFAC,
Geographical and phenotypic NPLS-DA,
Fluorescence [13,76]
discrimination of coffees UPLS-DA, MLR
and LDA
Predict moisture content, soluble
solids, total reducing sugars,
Prediction of coffee
lipids, proteins, trigonelline, NIR PLS [9,30,35,39,42,77–79]
chemical composition
5-CQA, caffeine, and sucrose
content of coffee
Predict CGA isomer composition
FTIR/MIR PLS [80,81]
and caffeine content
Determine trigonelline and
Fluorescence - [81,82]
caffeine content

Online monitoring of acidity during coffee roasting has also been studied using FT-
NIR techniques [38]. Coffee samples (arabica and robusta) were roasted using fast and
slow roasting designs at 220 ◦ C for 17 min and 183 ◦ C for 25 min, respectively. Online
acquisition of the spectra was done in the 10,000 and 4000 cm−1 range. In order to develop
a calibration model between titratable acidity and online-acquired NIR spectra, important
wavelength regions were selected (5150–5950 cm−1 , 4000–4760 cm−1 , 6320–7080 cm−1 , and
7470–9400 cm−1 ) and used in the development of a PLS model for acidity. These selected
regions are characteristic of 5-caffeoylquinic acid and were highlighted by Ribeiro et al. [40],
who aimed to link perceived acidity of beverages with arabica coffees chemical composition.
The proposed model had a RMSEP and range-error-ration (RER) of 0.16 and 11, respectively,
indicating its fairness in acidity prediction during coffee roasting. Yergenson and Aston [7]
carried out a similar study with the aim of determining coffee roast degree online by
controlling acidity. Coffee was roasted using four different temperature-time treatments
and NIR spectra acquired online in the 4000–7405 cm−1 range. Measurement of titratable
acidity over a varied range of roast times and temperatures followed by modeling the
findings allowed for the quantification of the relationships between degree of roast, acidity,
and roast time.
In another study, Santos et al. [31] explored the ability of FT-NIR spectroscopy in in-line
monitoring of coffee roasting process by measuring sucrose and colour. PLS regression was
utilized to develop predictive models by regressing spectral data against content of sucrose
and L*, a*, and b* colour values. For the case of sucrose, important wavenumbers with
significant contribution included 4980–5000, 5385–5410, 5325, 5075, and 5140 cm−1 . The
C=O second overtone region, first overtone and combination band region of O-H and C=O
bonds characterize the selected wavenumbers. Thus, the region could be associated with
carbohydrates, organic acids, and proteins. Colour characterization of coffee beans was
Processes 2022, 10, 71 11 of 25

based on L* parameters where wavenumbers 6100, 5300, 5200, and 4965 cm−1 were selected
as having the highest influence. The regions capture vibrations from C=O (5300 cm−1 ),
combination band vibrations from O-H (5200 cm−1 ), first overtone vibrations of C-H
(6100 cm−1 ), and combination vibrations of N-H (4965 cm−1 ). The authors concluded that
free amino acids, carbohydrates, and chlorogenic acids could be the main compounds
associated with colour during coffee roasting because they are involved in important
chemical reactions to form colour pigments. The developed PLS models showed a strong
prediction ability, with a co-efficient of determination of 0.85 and RER higher than 10.
Yergenson and Aston [6] explored the use of in-situ NIR spectroscopy in the prediction
of cracking events (start and end) during coffee roasting to provide a more vigorous
technique for controlling degree of roast based on the cracks. During coffee roasting, two
sets of popping sounds occur (first and second cracks) that are important roast degree
indicators for determining the roast endpoint. In this study, coffee samples were roasted
using different time-temperature profiles. Based on the PLS regression (PLSR) with audio
recordings from coffee roasting, in-situ NIR spectroscopy proved to be a reliable method
in predicting the start and end times of first and second crack events. Models exhibited a
good relationship between predicted and reference values with RMSEP values of 0.0068,
0.0091, 0.0041, and 0.0070 Au for the start of the first crack, end of the first crack, start of the
second crack, and end of the second crack, respectively.
Pires et al. [59] successfully employed multivariate calibration and NIR spectroscopy
as a substitute to the Agtron method to predict roast degrees in ground coffees and coffee
beans. With PLS the method, the development of mathematical models was based on
the relationship between data sets of NIR spectra and Agtron reference results to predict
Agtron values of new coffee samples. In order to build representative models, all profiles
of Agtron roasting were considered. The following NIR spectra regions were important for
modeling in the whole bean coffee model: 6770–6992, 7482–8000, 5774–5992, 4136–4391, and
4806–5257 cm−1 . For the ground coffee model, the NIR spectra regions included 7447–8000,
6783–6922, 5317–5445, 5034–5219, and 4173–4380 cm−1 . The developed models presented
promising results for roasting profiles prediction in roasted whole coffee beans as well as
ground coffees with RMSEP values of 4.48 and 3.67, respectively.
Real-time detection of faults during the coffee roasting process by NIR spectroscopy
and multivariate statistical process control (MSPC) based on principal component analysis
(PCA) has also been evaluated. In one study, five batches with enforced disturbances
(non-nominal batches) were prepared to mimic anomalous conditions of roasting and
to authenticate the detection capability of the MSPC method. The non-nominal batches
had deviations from nominal batches, i.e., higher amount of robusta coffee in the blend,
roasting with lower and higher roasting power, and use of higher and lower initial mass
of green coffee beans. PCA analysis of the acquired in-line NIR spectra demonstrated
clear differences among non-nominal batches and between nominal batches. The best
result was provided by a modelling method based on a time sliding window in terms of
differentiating batches with and without disturbances, resourcing to typical MSPC charts:
squared predicted error and Hotelling’s T2 statistics. In addition, a PCA model including
a four minute time window with three principal components efficiently detected all the
disturbances studied [60].

3.2. Prediction of Specialty Coffee Quality and Sensory Attributes


Sensory analysis of coffee is a complex process owing to the disparities of aromas,
flavours, as well as diversity of compounds produced during coffee roasting. The cupping
method is by far the most common professional technique for evaluating coffee, which
entails professional tasters evaluating coffee based on physical quality (size, odor, colour,
and shape of the beans) and cup quality (fragrance, aroma, astringency, acidity, flavour,
body, bitterness, aftertaste, and overall quality) [83]. However, this method is somewhat
subjective and requires professional personnel. Additionally, it is associated with various
shortcomings: it is time consuming, there is disparity in sensory perception amongst
Processes 2022, 10, 71 12 of 25

evaluators, a lack of reproducibility, and inadequate conclusions about samples, etc. In


view of this, spectroscopy techniques may be used as alternative tools in the sensory
evaluation of coffee as they are rapid and have the capability to reproduce results (Table 2).
Baqueta et al. [67] studied the ability of NIR spectroscopy in association with the
PLS method to determine coffees sensory characteristics. Coffee samples with different
variations including species, production region, variety, drying conditions, transport,
postharvest process, storage times, coffee blend, coffee composition, and roasting process
were used. Professional cuppers evaluated the samples based on the fragrance, acidity,
aroma, bitterness, body, flavour, astringency, overall quality, and aftertaste. NIR spectra of
the same samples were also taken. The performance of PLS models for each parameter were
validated using the following merit parameters: sensitivity, accuracy, linearity, residual
prediction deviation, fit, quantification, and detection limits. The developed models were
suitable to quantify, detect, differentiate, and predict sensory characteristics of coffee
samples, as all sensory characteristics were predicted with acceptable values consistent
with the merit parameters.
Similarly, Ribeiro et al. [40] carried out a study using NIR in conjunction with PLS with
the aim of developing predictive models for a reproducible and objective sensory analysis of
coffee. Additionally, the correlation between beverage sensory characteristics and chemical
composition of the green coffee bean was determined. Experts assessed the coffee samples
based on flavour, acidity, bitterness, body, overall quality, and cleanliness, assigning a score
to each attribute. PLS models were constructed for each attribute and they exhibited good
correlation between estimated values and those supplied by the experts with RMSEP values
of 0.37, 0.30, 0.25, 0.37, 0.42, and 0.30 for bitterness, acidity, flavour, cleanliness, overall
quality, and body, respectively. The researchers observed that, in roasted beans, lipids and
proteins were closely related to body attribute of the beverage, chlorogenic acids to acidity,
caffeine and chlorogenic acids to bitterness, and finally caffeine, trigonelline, chlorogenic
acid, polysaccharides, sucrose, and protein to flavour, cleanliness, and overall quality.
The market for specialty coffee is growing constantly due to the changing preferences
of consumers. Specialty coffees have a higher quality than commercial coffees. They
are characterized by unique flavours, have a known geographical origin, and commonly
comprise coffee beans certified as organic, or with rainforest alliance and fair-trade certi-
fications, etc. On the global market, specialty coffees are traded with a price premium of
approximately 20–50% when compared to regular coffees [84]. Therefore, there is a need
for more objective and reliable techniques for authenticating such coffees. Tolessa et al. [69]
developed a model to predict the quality of specialty coffee based on green beans using NIR
spectroscopy in conjunction with PLS analysis. Samples consisted of coffee from different
environments (low, mid, and high altitudes in Ethiopia) and processed differently (washed,
dry, and semi-washed methods). Professional coffee testers evaluated the samples based on
body, acidity, overall cup preference, aroma, uniformity, flavour, aftertaste, cup cleanness,
balance, and sweetness. The proposed models provided reliable predictions for the quality
of a specialty cup as well as its different quality characteristics with RMSEP values of 1.04,
0.27, 0.22, 0.27, and 0.24 for total quality, acidity, overall preference for the cup, aftertaste,
and body, respectively.
NIR spectroscopy has also been explored as an alternative to traditional methods of
sensory evaluation (cupping tests) to predict flavours of specialty coffees. Models were
developed using deep convolutional neural networks (DCNN) and the support vector
machine (SVM) method. Training was then done and used for the prediction of several
specialty coffee flavours (as presently being done by professional evaluators) utilizing
readings from NIR spectra as the inputs. Descriptions of flavour were classified into nine
flavour classes: fruity, floral, fermented/sour, vegetable/green, spices, roasted, sweet,
cocoa/nutty, and other. The two methods yielded comparable performance, having the
recall and accuracy of 70–73% for SVM and 75–77% for DCNN. This study offered an
innovative and objective method in the prediction of intricate flavours present in specialty
coffees [68]. A study by Arboleda [70] discriminated civet coffee from ordinary processed
Processes 2022, 10, 71 13 of 25

coffee utilizing NIR spectroscopy with the help of feed forward back propagation artificial
neural networks. The following wavelengths showed major differences among the two
sample groups: 1088 nm, 907 nm, 1650 nm, and 1540 nm. A good model was developed
with classification scores of 95–100%.
De-Araújo et al. [33] proposed a new non-destructive methodology based on NIR
spectroscopy and CACHAS (chemometrics-assisted colour histogram-based analytical
systems) for the verification of roasted-ground gourmet coffees (specialty coffee) with no
sample preparation. NIR spectra of the samples (36 traditional, 10 superior, and 44 gourmet)
were acquired in the 4000–10,000 cm−1 range. Digital images of the samples were made
using a scanner module from an HP Deskjet multifunctional printer. Histograms in the
hue-saturation-intensity (HSI), red-green-blue (RGB), and grayscale (GS) colour systems
from each sample were acquired from a circular region of interest (ROI). In this study,
one-class partial least squares (OC-PLS) and data-driven soft independent modeling of
class analogy (DD-SIMCA) were employed as one-class classifiers (OCC). Models were
developed using different pre-processing techniques and combinations of colour histograms
for NIR spectroscopy and CACHAS, respectively. Performance of the constructed models
were assessed with regard to sensitivity and specificity. DD-SIMCA using RGB histogram
for CACHAS and offset correction for the NIR spectra reported the best results as they both
correctly recognized 100% of the studied samples in the training set and test set.
The potential of MIR spectroscopy in association with chemometrics in predicting the
cup quality of coffees roasted to different roast degrees were examined (Table 2) [45]. Coffee
samples were first classified based on cup quality by skilled cuppers from worst to best.
The samples were then roasted to dark, medium, and light roast degrees. PCA was used to
show variability among the data as well as to detect probable outliers, whereas classification
models were developed using PLS discriminant analysis (PLS-DA). Different preprocessing
techniques of the raw data were compared and, with respect to PCA, the best clustering of
the samples was achieved with second derivative application to the wavenumber ranges of
1680–1800, 2800–2960, and 1130–1249 cm−1 . Models were developed based on a two-level
PLS-DA hierarchical strategy. In the first level, coffee was classified as of low or high quality
followed by separation based on quality of the cup in the second level. Spectral regions
2918, 1749, 1038, and 806 cm−1 were the most important for the differentiation between
coffees of low or high quality, whereas in the classification based on cup quality there was
no particular band region that characterized one particular class. However, there were
differences among the intensities. Similarly, Belchior et al. [8] investigated FTIR potential
for quantitative valuation of the sensory attributes of specialty coffees. Coffee samples were
evaluated by professional cuppers according to the Specialty Coffee Association of America
(SCAA) protocol for sensory analysis of coffee. Based on the scores, PLS regression models
were then constructed to establish and predict a sensory profile. The models showed a
good relationship between predicted and experimental data with low RMSEP values of 0.23.
However, there were no specific peaks significant for the coffee classification determination
when correlated to the SCAA scores for the entire spectra. This can be explained by the
fact that different compounds affecting the sensory profile of coffee absorb throughout the
entire spectrum.
Abreu et al. [44] compared the performance of ultraviolet-visible spectroscopy (UV-
Vis), high-performance liquid chromatography (HPLC), and MIR to discriminate between
traditional and specialty coffees using exploratory factor analysis (FA). Specialty and tra-
ditional roasted ground coffee samples were extracted using the best solvents (water and
dichloromethane), i.e., those that produced the most divergent metabolic fingerprints.
Among the methods investigated, HPLC and UV-Vis spectral fingerprints of the water
extracts provided more relevant discriminatory information than did MIR. Fatty acids,
chlorogenic acids, organic acids, trigonelline, and lipids were the most important com-
pounds for this discrimination. A method to discriminate espresso coffees based on their
sensory characteristics using attenuated total reflectance Fourier transform infrared spec-
troscopy (ATR-FTIR) and PLS-DA models was developed by Belchior et al. [71]. Samples
Processes 2022, 10, 71 14 of 25

were evaluated based on their flavour, aroma, acidity, aftertaste, and body, and models were
developed for each attribute. The models classified the samples based on these attributes
with good values of specificity and sensitivity for both calibration and validation sets.
Proteins, lipids, trigonelline, caffeine, carbohydrates, chlorogenic acids, and carboxylic
acids were linked to the body, flavour, and aftertaste. The acidity was linked to the presence
of chlorogenic acid, carboxylic acids, and alcohols, whereas esters, ketones, overall acids,
and aldehydes contributed to the aroma of the beverage.
Suhandy and Yulia [12] studied the feasibility of fluorescence spectroscopy in discrim-
inating Indonesian specialty coffees (Pea berry, Pagar Alam, and Civet). Soft independent
modeling of class analogy (SIMCA) and PCA were used as chemometric methods. The
PCA approach allowed grouping of the samples into three clusters where an excitation of
370 nm was associated with this observation. In contrast, SIMCA models showed a high
degree of specificity and sensitivity for calibration and prediction data sets.

3.3. Detection of Defective Beans and Adulterants in Coffee


Coffee has a high market value and as a result, its adulteration has become an exten-
sive practice for economic gain. It can include the addition of spent coffee grounds, chicory,
coffee husks, cereals (e.g., barley, corn, wheat, and rice) as well as the substitution of arabica
coffee, which is more expensive, with robusta coffee [85]. It is important to emphasize that
some countries have allowed the use of specific adulterants up to a certain level. However,
it is illegal if the information is not declared on the label. Therefore, fast and reliable
techniques are needed to protect consumers from fraudulent practices, and this has necessi-
tated development of new techniques based on spectroscopy (Table 2). Correia et al. [32]
proposed a methodology for identification and quantification of adulterants in roasted
coffee based on NIR spectroscopy and multivariate calibration by PLS and PCA. Green
coffee beans were roasted to different roast levels (dark, medium, and light), ground, and
mixed with adulterants (robusta coffee, corn, and peels/sticks) in a range of 1 to 100% w/w.
The PCA approach allowed for recognition of adulterated and pure samples with first and
second components accounting for over 90% of the total spectral variance. PLS models
with good sensitivity were constructed. However, the model for quantification of corn in
coffee samples was the best with RMSEP and coefficient of determination (R2 ) of 4% and
0.9788, respectively, compared to that of the peels/sticks. In regard to roast levels, light
roasted samples gave the best RMSEP values (2.8%).
Ebrahimi-Najafabadi et al. [61] studied the adulteration of ground-roasted coffee with
barley utilizing NIR spectroscopy and chemometrics methods. A widely applicable model
was constructed using nine varieties of coffee samples with their mixtures adulterated
with four types of barley, all, roasted to different roast degrees. Barley was mixed at levels
ranging from 2 to 20% w/w. A PLS regression was developed to obtain a quantitative model
for prediction of adulteration levels. The model proposed provided dependable predictions
of barley adulteration at levels as low as 2% w/w with low values of root mean square
errors (RMSE). Selection of variables was done using genetic algorithms, and the following
spectral regions were selected as the most significant in identifying coffee adulterated with
barley: 6032, 5748, 4788–4880, 4628–4688, 4336, and 4276 cm−1 . The regions are ascribed
to the first overtone of N–H, C=O, C–H, O–H, and S–H functional groups of ROH, ArOH,
H2 O, RNH2, and CONHR. In a similar approach, Winkler-Moser et al. [62] evaluated NIR
spectroscopy in conjunction with PLS regression analysis for corn adulteration detection
in ground-roasted coffee samples. Different samples were prepared by mixing corn with
coffee at concentrations of 0, 1, 5, 10, 15, and 20% w/w. Based on the confidence intervals
of predicted percent corn at specific levels of adulteration, the developed model seemed
not to be reliable in the detection of all coffee samples adulterated at levels below 4%. This
contrasts with the finding of Ebrahimi-Najafabadi et al. [61], whose model was able to
detect coffee adulteration with barley at 2% levels.
The capability of NIR hyperspectral imaging in combination with multivariate curve
resolution-alternating least squares (MCR-ALS) to detect and quantify adulterants in
Processes 2022, 10, 71 15 of 25

roasted-ground coffee has also been reported [86]. Four coffee adulterants were studied,
i.e., coffee husks, wood sticks, soil, and corn kernels. They were added to the roasted-
ground coffee in levels of 1.0 to 40% (w/w). Quantification of adulterants was based on
the utilization of the MCR correlation constraint, whereas detection was achieved through
comparison of the reference spectrum of each adulterant with the spectra recovered by
the MCR model utilizing only the adulterated samples. For detection, the values of the
coefficient of correlation of the studied adulterants ranged from 0.90 to 0.99. However, the
absolute errors for the quantification of adulterants in the developed models were below
4%. The findings pointed out the practicability in the application of the methodologies
proposed. Flores-Valdez et al. [47] successfully employed FT-NIR spectroscopy coupled
with chemometrics for identification and quantification of adulterants in coffee. Coffee
samples were adulterated with soy, corn, barley, oat, coffee husks, and rice, in proportions
ranging from 1 to 30% in increments of 1%. They developed a SIMCA model based
on the percentage rejection (selectivity) and recognition (sensitivity), which allowed the
discrimination of adulterated and unadulterated samples with an accuracy of 100%. The
quantitative model with the highest performance corresponded to the PLS1 algorithm
with an R2 ≥ 0.99, standard error of prediction (SEP) of 0.45–0.94, and standard error of
calibration (SEC) of 0.39–0.82. To build this model, the most significant spectral range
considered were 800–1800 and 2800–3500 cm−1 because they showed the highest variability
of the spectra, signifying the change in absorbance as well as increase in the levels of
adulterants in the samples. In another study, Brondi et al. [63] compared differential
scanning calorimetry (DSC) and attenuated total reflectance-Fourier transform infrared
(ATR-FTIR) spectroscopy methodologies together with chemometric analysis for detection
and quantification of corn in roasted-ground coffee. Adulterated samples were prepared
mixing coffee roasted to light, medium, and dark roasts with ground roast corn in levels
ranging from 0.5 to 40% (m/m). PLS and PCA models were constructed to quantify
adulteration levels and to distinguish non-adulterated samples from adulterated ones,
respectively. PCA models built with FTIR and DSC data were able to detect adulteration
of coffee with corn in concentrations below 1%. Conversely, PLS models demonstrated a
good correlation between reference concentrations and estimated values with RMSEP of
1.53% for FTIR and 3.94% for DSC.
The presence of defective beans (i.e., insect damaged, broken, sour, black, mouldy, or
immature) significantly affects the price and quality of the batch as well as the beverage.
According to Franca et al. [87], defective beans roast differently, acquiring a different colour
after roasting and imparting undesirable organoleptic characteristics to the beverage, which
decreases cup quality. Therefore, it is vital to develop techniques for rapid and accurate
assessment of coffee quality. Craig et al. [41] successfully developed a technique based
on FTIR and NIR to quantify defects in roasted coffees. Quantitative models based on
PLS regression were developed with the aim of predicting defective bean percentage in
mixtures with non-defective beans. The mixtures were prepared by mixing each type
of defective bean (immature, black, light sour, and dark sour) with non-defective ones,
with the defect amounts ranging from 3 to 30% in steps of 3%. PLS regression models
provided accurate predictive results with correlation coefficients and root mean square
error of validation (RMSEV) of 0.891 and 0.032, respectively, for FTIR, and 0.953 and 0.026
for the NIR technique. Comparing the two methods, quantitative models from NIR were
considerably more robust as compared to the ones based on FTIR for quantifying defective
beans in roasted coffees. With the employment of a new statistical approach, Elastic net, the
same authors evaluated the performance of NIR and ATR-FTIR for the discrimination of
roasted defective (immature, black, dark sour, and light sour) and non-defective coffees. The
developed models based on Elastic net allowed for the correct classification of non-defective
and defective coffees with amino acids/proteins, carbohydrates, lipids, chlorogenic acids,
and caffeine being the chief chemical descriptors that characterized the samples. Generally,
both NIR and ATR-FTIR proved to be reliable methods for the discrimination of roasted
defective and non-defective coffees.
Processes 2022, 10, 71 16 of 25

In another study, Santos et al. [64] assessed the ability of NIR spectroscopy in discrimi-
nating defective beans (sour, immature, and black) from non-defective ones using PCA and
PLSR analysis. Although PCA modelling did not allow the separation of all the studied
defects, PLS regression linking the NIR spectrum to the mass fraction of non-defective and
defective beans showed relative errors of about 5%. This is an indication that with the
proposed methodology, it may be possible to do quantification of non-defective quality
against other qualities in a batch. Chen et al. [65] studied the capability of VISIBLE-NEAR
INFRARED (Vis-NIR) hyperspectral imaging for insect-damaged beans detection. The
experimental samples included 570 and 569 defective and healthy beans, respectively. A
Vis-NIR push-broom hyperspectral imaging camera was used to attain the beans images,
a hyperspectral insect damage detection algorithm (HIDDA) was used for selection of
bands, and a convolutional neural network (CNN) and constrained energy minimization-
support vector machine (CEM-SVM) was used for identification. Classification accuracy
of defective beans was 93% and that of healthy beans was 96.4%. It was observed that the
wavelength range of 850–950 nm was important in accurately identifying insect damaged
and healthy beans.
Application of diffuse reflectance (DR) and attenuated total reflectance (ATR) Fourier
transform infrared spectroscopy for discrimination of mature and immature green coffees
has also been studied. Agglomerative hierarchical clustering (AHC) and PCA analysis of
the normalized ATR and DR spectra clearly separated the samples into two groups, mature
and immature. Spectral ranges that significantly contributed to sample clustering were as
follows 2074–3100, 1800–1995, 1477–800, and 1725–1762 cm−1 for DR spectra and 3030–3080
and 1533–1535 cm−1 for ATR spectra. Linear discriminant analysis (LDA) classification
models based on the absorbance readings after normalization at four wavenumbers (1141,
1741, 2852, and 2921 cm−1 ) were developed. The results provided by the models presented
recognition as well as prediction capabilities of 100%, for both ATR and DR data [49]. The
same research group also employed similar techniques with an aim of characterizing and
discriminating between non-defective and defective (sour, immature, black) coffee beans
prior to roasting. Similar observations were reported with PCA and AHC of the reflectance
spectra showing discrimination between non-defective and defective beans [66].
Another study focused on the application of Fourier transform infrared-photoacoustic
spectroscopy (FTIR-PAS) in a quantitative assessment of defects in roasted ground cof-
fee [51]. Mixtures of different defects (broken, sour, black, peels, and woods) were mixed
with non-defective coffee of robusta and arabica species in precise ratios to form different
classes of blends. All samples were medium roasted and ground. With PCA, it was possible
to predict the fraction/amount as well as the defects nature in blends. PLS-DA provided
information on the similarities between the blends. The developed model classified samples
into four classes with 100% accuracy and specificities of higher than 0.9.

3.4. Discrimination of Coffee Based on Species, Variety, and Geographical Origin


Coffee identification has become an important aspect of the global coffee market
considering the great variability in quality and selling price based on its geographic origin,
species, and variety. This has necessitated the development of fast and reliable methods
for discrimination of coffee based on the aforementioned parameters to avoid fraudulence.
Arabica coffees are usually considered to be of better quality than robusta coffees in
terms of flavour and taste. NIR spectroscopy in combination with the direct orthogonal
signal correction (DOSC) preprocessing technique was utilized to develop an analytical
technique to discriminate between pure robusta and arabica species and their blends.
Classification models from raw and corrected NIR spectra were developed using the
potential function class-modelling technique. It was concluded that the combination of
DOSC, NIR spectroscopy, and the potential functions method could be utilized as an
ideal approach to distinguish between pure robusta and arabica coffee varieties and to
differentiate between blends and pure varieties of the two species [72]. In another study,
Zhang et al. [48] sought to establish the best pattern recognition method that could be
Processes 2022, 10, 71 17 of 25

employed in conjunction with MIR spectroscopy for coffee variety identification. PCA
successfully categorized each coffee variety with the first four principal components (PCs),
explaining over 99% of the total variance. Twenty-nine best wavenumbers were selected
by the loadings of the first four PCs and used in the classification models development
utilizing ten different pattern recognition techniques. The extreme learning machine (ELM),
radial basis function neural network (RBFNN), back propagation neural network (BPNN),
relevance vector machine (RVM), SIMCA, and SVM models were categorized as highly
effective techniques, with classification accuracies of more than 95% in the test and training
set. In contrast, naive Bayes classifier and PLS-DA, K-nearest neighbors, and random
forest were categorized as techniques of low effectiveness and medium effectiveness with
classification accuracy below 80% and above 80% in the test and training, respectively.
Marquetti et al. [37] evaluated the potential of NIR spectroscopy and PLS-DA in dis-
tinguishing green arabica coffee samples based on their geographical origin and genotypic
characteristics. Coffee samples of four genotypes (IA 59, IPR 106, IPR 99, and IPR 105)
and from four different cities were studied. The best PLS-DA models were developed
from the pretreated spectra (Savitzky–Golay second derivative and multiplicative scatter
correction) and were capable of distinguishing coffee samples both genotypically and geo-
graphically. However, the specificity and sensitivity for prediction and calibration sets for
geographic discrimination were greater than for the genotypic one. The best model for both
parameters correctly identified 94.4% of validation samples. Similarly, Giraudo et al. [36]
used NIR spectroscopy and multi-variate data analysis to classify coffee cultivated in
different countries and continents. Interval PLS-DA and laboratory-independent partial
least square-discriminant analysis models were developed by following a ranked method,
i.e., by first putting into consideration the continent followed by the country of origin as
the rule of discrimination. For the continent and country of origin, the best classification
models correctly identified over 98% and 100% of validation samples, respectively.
Medina et al. [73] compared the capability of nuclear magnetic resonance (1 H-NMR),
NIR, and ATR-MIR spectroscopic methods in discriminating coffee based on its species
(arabica and robusta) and geographic origin (Colombian coffee versus other origins). Con-
sidering species, samples were successfully discriminated as expected by all the techniques
using the built PLS classification models. However, for the case of origin determination,
ATR-MIR and 1 H-NMR exhibited comparable capability to discriminate Colombian coffee
samples, but weak results were observed with the NIR spectroscopy technique. In contrast,
Bona et al. [52] evaluated the ability of NIR and FTIR (MIR) spectroscopy in conjunction
with support vector machines (SVM) for geographical classification of different arabica
coffee genotypes and reported better results for the NIR method. The technique exhibited
greater performance as compared to FTIR, with a specificity and sensitivity of 100% in
discriminating coffee samples from different regions. Okubo and Kurata [34] proposed an
approach for geographical classification of different species of coffee (arabica and robusta)
exploiting NIR spectroscopy and SIMCA as the classification methods. Models were devel-
oped and their distance from SIMCA evaluated to establish if they were different. Although
some samples were partially classified into several categories, a good classification result
was obtained with over 73% correct classification rate. Model distance values from SIMCA
were relatively similar among samples, which could explain the misclassification.
Wongsaipun et al. [5] applied an artificial neural network technique (a self-organizing
map (SOM)) with the aim of tracing geographical origin of arabica coffee beans grown in
three provinces in northern Thailand using NIR spectroscopy. The three provinces included
Chiang Mai, Lampang, and Mae Hong Son. The self-organizing map discrimination index
(SOMDI) was applied to identify important parameters of the spectroscopic data. SOM
analysis of the NIR spectra separated the samples from the three different regions. Based on
the SOMDI results, the coffee samples from Chiang Mai could be well discriminated using
the following NIR spectral ranges: 2260–2498, 1896–2180, 1254–1326, and 880–1182 nm.
Luna et al. [74] evaluated different chemometric methods for the classification of robusta
coffee cultivars using FT-NIR spectroscopy. Five coffee cultivars were studied. SOM,
Processes 2022, 10, 71 18 of 25

SIMCA, PLS-DA, and SVM were explored for the cultivars classification, whereas PCA
was used for identification of the clusters. PCA analysis of the pre-processed spectra
discriminated samples into different groups, which were correlated to the presence of water,
caffeine, lipids, sugars, chlorogenic acids, carbohydrates, and proteins. SOM presented
the best results, providing 100% correct identification of the validation samples, whereas
SIMCA, PLS-DA, SVM using 4 PCs and SVM using 3 PCs provided 99.6, 82.9, 99.6, and
82.9%, respectively.
A study by Link et al. [50] distinguished coffee samples of different geographical
origins and genetic make-up using FTIR spectroscopy in tandem with radial-basis function
networks (RBFs). Optimization of some parameters was done using sequential simplex
in order to select the best neural network. The optimized RBFs correctly classified the
samples both genotypically (94.44%) and geographically (100%). This chemometric method
exhibited superior performance compared to multilayer perceptron (MLP) and SIMCA
developed for classification of coffee as reported by the same research group. A feasibility
study using ATR-FTIR spectroscopy and chemometric analysis with the aim of discrim-
inating coffees from different topographical origins as well as of different roast degrees
was conducted by Wang et al. [75]. Coffee samples from different origins (Kenya, Colom-
bia, Ethiopia, and Costa Rica) were roasted to two roast degrees (medium and dark) and
extracted using six organic solvents (hexane, ethyl acetate, dichloromethane, acetic acid,
ethanol, and acetone) and a mixture of an equal volume of the solvent and water. Extrac-
tion using water and the solvent resulted in clean extracts, which provided good spectral
information important for sample discrimination. Developed classification models based
on SIMCA and of dark roasted coffee correctly classified samples (100%) based on their
origin when ethyl acetate solvent was used.
The potential of Raman spectroscopy in coffee species and cultivar discrimination has
been investigated. A study by Rubayiza and Meurens [53] discriminated arabica, liberica,
and robusta coffees based on their lipid fraction using Fourier transform (FT) Raman
spectroscopy. Raman measurements of the lipid fraction from roasted and green coffee
beans was taken using a laser emissions at a wavelength of 1064 nm and resolution of
4 cm−1 . PCA analysis of the spectra revealed clustering of the samples based on the species
with the first principal component (PC1), explaining 93% of the spectral variance. This was
associated with the concentration of kahweol, which manifests at two specific scattering
bands at 1478 and 1567 cm−1 . Keidel et al. [54] reported similar findings after analyzing
green coffee beans belonging to robusta and arabica species from different geographical
origins (South America, Asia, and Africa). Measurements were taken using a 1064 nm laser
and a spectral resolution of 4 cm−1 . It was possible to classify coffee samples based on the
specific kahweol content. In a similar approach, Wermelinger et al. [55] evaluated Raman
spectroscopy for quantification of robusta coffee fraction in blends by analyzing the lipid
fraction. Pure robusta and arabica and their mixtures with contents of robusta of 5, 10, 25,
33, 50, and 75 wt.% were used. Lipids were extracted from roasted coffee beans and their
Raman spectra collected at 532 nm. It was possible to determine the robusta content in
studied mixtures by examining the intensity ratio between the Raman peaks of fatty acids
at 1665 or 1460 cm−1 and the kahweol peak at 1570 cm−1 .
Luna et al. [11] compared different chemometric approaches that could be used to
classify robusta coffee cultivars using Raman spectroscopy. Raman spectra of five coffee
cultivars were collected at 785 nm with a laser excitation power of 100 mW. The follow-
ing classification methods were studied: quadratic discriminant analysis (QDA), mixture
discriminant analysis (MDA), PLS–DA with Bayesian inference, SIMCA, regularized dis-
criminant analysis (RDA), and LDA. Two preprocessing methods, multiplicative scatter
correction (MSC) and mean centering (MC), were also compared, with MSC yielding more
accurate results for all the classification methods. Using multiplicative scatter correction,
RDA, MDA, QDA, SIMCA, and PLS–DA methods correctly classified 100% of the samples,
whereas LDA correctly classified 98.7% of the samples. Conversely, using MC, correct
classification of the samples was 70.7%, 62.7%, 62.7%, 62.7%, 61.3%, and 97.3% for MDA,
Processes 2022, 10, 71 19 of 25

LDA, RDA, QDA, PLS-DA, and SIMCA, respectively. A study by Figueiredo et al. [56]
differentiated green coffee beans of different genotypes using Raman spectroscopy in con-
junction with chemometric methods (PLS-DA and PCA). Acquisition of the Raman spectra
were made using an excitation laser with λ = 1064 nm at a spectral resolution of 4 cm−1 . It
was possible to distinguish the coffee genotypes based on spectral bands that are typical
of kahweol (1567 and 1479 cm−1 ) and fatty acids (1442 and 1302 cm−1 ) using PCA. The
developed PLS-DA models demonstrated a good association between the training and test
set of the studied samples indicating its suitability in performing the task.
El-Abassy et al. [14] proposed a method capable of discriminating between robusta
and arabica coffee species based on their lipid and chlorogenic acid (CGA) contents using
Raman spectroscopy. Measurements of the samples were made using a 514.5 nm laser
with an excitation power of 10 mW. PCA analysis of the whole Raman spectra presented
a clear discrimination between robusta and arabica coffee with 93% of the total spectral
variation. The most important spectral range that contributed to this observation was
1000–1750 cm−1 , an area dominated with Raman bands of CGA. Discrimination of the two
coffee species was also observed when PCA analysis was restricted to the spectral range
between 2700–3050 cm−1 , which is dominated by bands of lipids. In this respect, the first
two PCs constituted 85% of explained variation in the spectra. Studies on fluorescence
spectroscopy as a valuable technique to discriminate coffee based on their geographical
origin and phenotypic characteristics are also present in the literature. Botelho et al. [13]
developed a supervised classification technique using chemometrics tools with the ability
to discriminate coffee of different origins (four) using fluorescence spectroscopy. Discrim-
ination models were constructed by employing N-way partial least squares (NPLS-DA),
parallel factor analysis (PARAFAC), and unfolded partial least squares (UPLS-DA) methods.
Among the three methods studied, UPLS-DA models showed the best results, with f-scores
of coffees from two regions above 0.8, for both test and training sets. In another study,
Dankowska et al. [76] compared the ability of UV-Vis spectroscopies, fluorescence, and the
low- and mid-level data fusion of both spectroscopies to quantify concentrations of roasted
robusta and arabica in coffee blends. Classification models were developed using multiple
linear regression (MLR) and LDA. The best prediction capability of the MLR models was
achieved with application of the mid-level data fusion model of fluorescence and UV-Vis
intensities at 60 nm wavelength interval with RMSEV of 7.9% and RMSEC of 3.6%. Better
discrimination ability was obtained with data fusion with the highest classification accuracy
(above 96.0%) being achieved for the low-level LDA model with fluorescence intensities at
a 60 nm wavelength interval.

3.5. Prediction of Coffee Chemical Composition


Coffees chemical composition affects the quality of the beverage, making its analysis a
necessity for coffee processors. The use of known conventional methods to determine these
compounds in coffee is laborious and expensive due to sample preparation. Spectroscopic
analysis of ground or whole coffee beans is a potential alternative to overcome these limita-
tions. Studies on the application of NIR spectroscopy for the quantification of the chemical
components of green coffee beans have been done (Table 2). Tugnolo et al. [30] compared
the potential of NIR spectroscopy and thermogravimetric analysis in measuring moisture
content of roasted beans and their grounds. The developed PLS models from NIR data
demonstrated a good relationship between estimated values and reference concentrations
with errors in prediction below 0.15% and coefficients of determination higher than 0.95. To
compare the measurements of the two techniques, the passing-Bablok regression method
was performed. Even though there were significant differences, the proposed residual
dispersion index (RDI%) showed higher predictive accuracy of NIR-based predictions
(RDI% = 5.93) with regard to thermogravimetric analysis measurements (RDI% = 9.68).
Escobar et al. [77] developed a rapid method for moisture content prediction in unroasted
coffee beans using NIR spectroscopy. Among the samples analyzed (two accessions of
liberica coffee and four varieties of arabica coffee), the calibration model for typica coffee
Processes 2022, 10, 71 20 of 25

(arabica variety) exhibited the best performance in predicting the moisture content of stud-
ied samples as evidenced by its high cross-validated coefficient of determination (0.8556)
and cross-validated root mean square error of prediction (0.6355).
Kyaw et al. [78] reported promising results in relation to the moisture content pre-
diction of ground unroasted coffee beans using NIR spectroscopy. Good accuracy for the
prediction of moisture content was obtained from the spectral data pretreated with second
derivative and Kubelka-Munk (K/S) data transformation (correlation coefficient (r) = 0.87
and accuracy = 99%). Zhu et al. [35] proposed a rapid method for lipid and protein content
determination in unroasted coffee beans using NIR spectroscopy and chemometrics. Or-
thogonal signal correction (OSC) and different spectral pretreatment methods (standard
normal variate (SNV), MSC, 1st or 2nd derivative, and Savitzky-Golay smoothing) were
compared during the process of building a PLS regression model. Model quality was en-
hanced by the MSC, SNV, and OSC pretreatment methods. On the contrary, the 1st and 2nd
derivative reduced the model quality. Important variable selection significantly enhanced
the PLS models, with OSC-PLS models standing out as the most robust for prediction of
lipids and proteins with an RMSEP of less than 0.106 and a coefficient of determination
for prediction of more than 0.982. In a similar approach, Macedo et al. [79] proposed a
methodology based on the use of NIR spectroscopy associated with PLS regression to
estimate some chemical properties (soluble solids, total and reducing sugars, and moisture
content) in intact green coffee samples. The highest R2 obtained for the samples in the
validation set included 0.781, 0.810, 0.694, and 0.516 for reducing sugars, moisture content,
total sugar, and soluble solids, respectively.
In another study, NIR spectroscopy coupled to chemometrics (PLS regression) was
investigated to predict 5-caffeoylquinic acid (5-CQA), trigonelline, and caffeine content
in green coffee beans. The prediction models for 5-CQA, trigonelline, and caffeine were
constructed using 5, 6, and 7 latent variables, giving RMSEPs of 0.27, 0.07, and 0.08 and
correlation coefficients of cross validation (Rcv ) of 0.96, 0.96, and 0.98, respectively [39]. An
analytical technique based on NIR spectroscopy for prediction of caffeine, proteins, lipids,
CGA, phenolic compounds, sucrose, and total sugars was investigated in grounds of green
coffee beans [42]. A modified PLS regression was employed in the development of pre-
diction models. Prediction models for lipids (ratio of performance deviation (RPD) = 2.77,
R2 = 0.87), phenolic compounds (RPD = 2.62, R2 = 0.86), total sugars (RPD = 2.55, R2 = 0.85),
and sucrose (RPD = 2.44, R2 = 0.84) were not very accurate. Models for proteins (RPD = 4.09,
R2 = 0.94), caffeine (RPD = 4.16, R2 = 0.92), and chlorogenic acids (RPD = 4.16, R2 = 0.94),
presented the best prediction capabilities. Likewise, Safrizal et al. [9] reported calibration
models for the prediction of caffeine, lipid, and CGA content with an r-value and RPD
above 0.7 and 2, respectively.
The literature also contains investigations on the feasibility of FTIR/MIR spectroscopy
in the prediction of chemical components in coffee. Liang et al. [80] used ATR-FTIR
spectroscopy together with PLS regression with the aim of developing a rapid method
to determine antioxidant activity and CGA isomer composition in roasted and unroasted
coffee beans. FTIR spectral data were processed (i.e., baseline corrected and mean centered)
prior to model development. For the case of CGA isomer concentrations, the models
presented a good relationship between HPLC reference values and estimated values with
Rcv of more than 0.92 for all the isomers. Good models were also constructed with the ability
to predict antioxidant capacities in coffee beans. In another study, Weldegebreal et al. [81]
successfully developed rapid methods for direct caffeine content determination in aqueous
solution of unroasted coffee beans using fluorescence, NIR, and ATR-FTIR spectroscopy.
Caffeine content of the samples was determined using the three types of equipment and
the results were compared to the standard method (UV-Vis spectroscopy) for validation.
The results for the newly developed techniques were similar to the results achieved by the
standard method. In a similar approach, Yisak et al. [82] developed a reliable and sensitive
technique for simultaneous trigonelline and caffeine determination in the aqueous extract
of unroasted coffee beans with relative standard deviations below 4%. Evaluation of the
Processes 2022, 10, 71 21 of 25

developed analytical method was done by spiking green coffee beans with trigonelline and
caffeine standards. Average recoveries of 99 ± 2% for both the alkaloids was reported.

4. Future Perspectives and Conclusions


The information presented in this review clearly confirms the potential of spectroscopic
techniques in coffee assessment. However, most of these studies are purely academic.
Mendes and Duarte [23] highlighted issues that must be addressed for the developed
spectroscopic analytical methods to be applied as official methods of analysis in the coffee
industry. They include guidelines on the validation of the methods, use of samples from
credible sources (known origin) for method development, use of a satisfactory number
of representative samples in order to cover all sample variances for development of a
robust model, and choosing the most appropriate chemometric technique for a particular
objective. The findings of this review demonstrate the potential of spectroscopic methods
to evaluate coffee for different parameters. However, there are other aspects that need
further investigation, including prediction of coffees processed using different methods,
differentiation of organic from inorganic coffee, and monitoring of coffee quality during
storage, of which limited scientific information is available. Additionally, Raman and
fluorescence spectroscopy received very little attention, with most applications being
focused on NIR spectroscopy. Available information in the field of coffee spectroscopy
primarily focused on IR spectroscopy. To broaden this scope, this review includes recent
studies based on IR, Raman, and fluorescence spectroscopy.
In conclusion, coffee composition and quality assessment are an important practice
in the global coffee trade market. Conventional methods that are most frequently used to
perform these tasks are often time-consuming, complex, expensive, and require several
steps in sample preparation. Spectroscopic techniques (NIR, MIR/FTIR, Raman, and
fluorescence), as evidenced by the reviewed literature, could be used to evaluate coffee
for different parameters along the production process and in the near future replace the
conventional methods. The vast majority of applications are based on NIR spectroscopy;
only a few examples with Raman and fluorescence spectroscopy are available in the
literature. In the future, applications using these methods will likely become more common,
as valuable information can also be obtained from these spectra. The major advantages of
these techniques are that they are rapid, require no or simple sample preparation, are non-
destructive, and can be easily integrated into processes. Nevertheless, it is important for
laboratory studies to upscale to industrial applications if these advantages are to be realized.

Author Contributions: Conceptualization, writing—review and editing, B.H.; writing—original


draft preparation, L.M.; writing—review and editing, D.N. All authors have read and agreed to the
published version of the manuscript.
Funding: This research was funded by the German Academic Exchange Service (DAAD), funding ID
57524989. However, there was no funding of the publication costs.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: No new data were created or analyzed in this study. Data sharing is
not applicable to this article.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. ICO Trade Statistcis. Available online: http://www.ico.org/trade_statistics.asp?section=Statistics (accessed on 31 August 2021).
2. Belitz, H.-D.; Grosch, W.; Schieberle, P. Coffee, tea, cocoa. In Food Chemistry, 4th ed.; Belitz, H.-D., Grosch, W., Schieberle, P., Eds.;
Springer: Berlin, Germany, 2009; pp. 938–951.
3. Yashin, A.; Yashin, Y.; Xia, X.; Nemzer, B. Chromatographic Methods for Coffee Analysis: A Review. J. Food Res. 2017, 6, 60.
[CrossRef]
Processes 2022, 10, 71 22 of 25

4. Alonso-Salces, R.M.; Serra, F.; Reniero, F.; Héberger, K. Coffea arabica and Coffea canephora): Chemometric evaluation of
phenolic and methylxanthine contents. J. Agric. Food Chem. 2009, 57, 4224–4235. [CrossRef] [PubMed]
5. Wongsaipun, S.; Theanjumpol, P.; Muenmanee, N.; Boonyakiat, D.; Funsueb, S.; Kittiwachana, S. Application of Artificial Neural
Network for Tracing the Geographical Origins of Coffee Bean in Northern Areas of Thailand Using Near Infrared Spectroscopy.
Chiang Mai J. Sci. 2021, 48, 163–175.
6. Yergenson, N.; Aston, D.E. Monitoring coffee roasting cracks and predicting with in situ near-infrared spectroscopy. J. Food
Process Eng. 2020, 43, 13305. [CrossRef]
7. Yergenson, N.; Aston, D.E. Online determination of coffee roast degree toward controlling acidity. J. Near Infrared Spectrosc. 2020,
28, 175–185. [CrossRef]
8. Belchior, V.; Botelho, B.G.; Casal, S.; Oliveira, L.S.; Franca, A.S. FTIR and Chemometrics as Effective Tools in Predicting the Quality
of Specialty Coffees. Food Anal. Methods 2020, 13, 275–283. [CrossRef]
9. Safrizal; Sutrisno; Pujantoro, L.E.N.; Ahmad, U. Samsudin Predicting Lipid, Caffeine and Chlorogenic Acid Contents of Arabica
Coffee Using NIRS. Int. J. Sci. Technol. Res. 2019, 8, 367–370.
10. Reis, N.; Botelho, B.G.; Franca, A.S.; Oliveira, L. Simultaneous Detection of Multiple Adulterants in Ground Roasted Coffee by
ATR-FTIR Spectroscopy and Data Fusion. Food Anal. Methods 2017, 10, 2700–2709. [CrossRef]
11. Luna, A.S.; DaSilva, A.P.; DaSilva, C.S.; Lima, I.C.A.; DeGois, J.S. Chemometric methods for classification of clonal varieties of
green coffee using Raman spectroscopy and direct sample analysis. J. Food Compos. Anal. 2019, 76, 44–50. [CrossRef]
12. Suhandy, D.; Yulia, M. Discrimination of several Indonesian specialty coffees using Fluorescence Spectroscopy combined with
SIMCA method. IOP Conf. Ser. Mater. Sci. Eng. 2018, 334, 012059. [CrossRef]
13. Botelho, B.G.; Oliveira, L.S.; Franca, A.S. Fluorescence spectroscopy as tool for the geographical discrimination of coffees produced
in different regions of Minas Gerais State in Brazil. Food Control 2017, 77, 25–31. [CrossRef]
14. El-Abassy, R.M.; Donfack, P.; Materny, A. Discrimination between Arabica and Robusta green coffee using visible micro Raman
spectroscopy and chemometric analysis. Food Chem. 2011, 126, 1443–1448. [CrossRef]
15. Zettel, V.; Ahmad, M.H.; Beltramo, T.; Hermannseder, B.; Hitzemann, A.; Nache, M.; Paquet-Durand, O.; Schöck, T.; Hecker, F.;
Hitzmann, B. Supervision of food manufacturing processes using optical process analyzers—An overview. ChemBioEng Rev. 2016,
3, 219–228. [CrossRef]
16. Ahmad, M.H.; Sahar, A.; Hitzmann, B. Fluorescence spectroscopy for the monitoring of food processes. Adv. Biochem. Eng.
Biotechnol. 2017, 161, 121–152.
17. Nawrocka, A.; Lamorska, J. Determination of Food Quality by Using Spectroscopic Methods. In Advances in Agrophysical Research;
Grundas, S., Stepniewski, A., Eds.; Intechopen: London, UK, 2013; pp. 348–363.
18. Lin, M.; Rasco, B.A.; Cavinato, A.G.; Al-Holy, M. Infrared (IR) Spectroscopy—Near Infrared Spectroscopy and Mid-Infrared
Spectroscopy. In Infrared Spectroscopy for Food Quality Analysis and Control; Sun, D.-W., Ed.; Academic Press: London, UK, 2009;
pp. 119–143.
19. Dufour, E. Principles of Infrared spectroscopy. In Infrared Spectroscopy for Food Quality Analysis and Control; Sun, C., Ed.; Academic
Press: London, UK, 2009; pp. 1–27.
20. Prieto, N.; Pawluczyk, O.; Dugan, M.E.R.; Aalhus, J.L. A Review of the Principles and Applications of Near-Infrared Spectroscopy
to Characterize Meat, Fat, and Meat Products. Appl. Spectrosc. 2017, 71, 1403–1426. [CrossRef]
21. Reich, G. Near-infrared spectroscopy and imaging: Basic principles and pharmaceutical applications. Adv. Drug Deliv. Rev. 2005,
57, 1109–1143. [CrossRef]
22. Nelson, D.L. Spectroscopic Methods in Food Analysis, 1st ed.; Franca, A.S., Nollet, L., Eds.; CRC Press: Boca Raton, FL, USA, 2018.
23. Mendes, E.; Duarte, N. Mid-Infrared Spectroscopy as a Valuable Tool to Tackle Food Analysis: A Literature Review on Coffee,
Dairies, Honey, Olive Oil and Wine. Foods 2021, 10, 477. [CrossRef]
24. Stehle, C.U.; Abuillan, W.; Gompf, B.; Dressel, M. Far-infrared spectroscopy on free-standing protein films under defined
temperature and hydration control. J. Chem. Phys. 2012, 136, 075102. [CrossRef] [PubMed]
25. Wang, Q. Raman Spectroscopic Characterization and Analysis of Agricultural and Biological Systems. 2013. Available online:
https://dr.lib.iastate.edu/handle/20.500.12876/27208 (accessed on 12 October 2021).
26. McCreery, R.L. Calibration and Validation in Raman Spectroscopy for Chemical Analysis; Wiley-Interscience: New York, NY, USA, 2000.
27. Petersen, M.; Yu, Z.; Lu, X. Application of Raman Spectroscopic Methods in Food Safety: A Review. Biosensors 2021, 11, 187.
[CrossRef] [PubMed]
28. Boyaci, I.H.; Temiz, H.T.; Genis, H.E.; Soykut, E.A.; Yazgan, N.N.; Guven, B.; Uysal, R.S.; Bozkurt, A.G.; Ilaslan, K.; Toruna, O.; et al.
Dispersive and FT-Raman spectroscopic methods in food analysis. R. Soc. Chem. 2015, 5, 56606–56624. [CrossRef]
29. Khuwijitjaru, P.; Boonyapisomparn, K.; Huck, C.W. Near-infrared spectroscopy with linear discriminant analysis for green
‘Robusta’ coffee bean sorting. Int. Food Res. J. 2020, 27, 287–294.
30. Tugnolo, A.; Giovenzana, V.; Malegori, C.; Oliveri, P.; Casson, A.; Curatitoli, M.; Guidetti, R.; Beghi, R. A reliable tool based
on near-infrared spectroscopy for the monitoring of moisture content in roasted and ground coffee: A comparative study with
thermogravimetric analysis. Food Control 2021, 130, 108312. [CrossRef]
31. Santos, J.R.; Viegas, O.; Páscoa, R.N.M.J.; Ferreira, I.M.P.L.V.O.; Rangel, A.O.S.S.; Lopes, J.A. In-line monitoring of the coffee
roasting process with near infrared spectroscopy: Measurement of sucrose and colour. Food Chem. 2016, 208, 103–110. [CrossRef]
Processes 2022, 10, 71 23 of 25

32. Correia, R.M.; Tosato, F.; Domingos, E.; Rodrigues, R.R.T.; Aquino, L.F.M.; Filgueiras, P.R.; Lacerda, V.; Romão, W. Portable near
infrared spectroscopy applied to quality control of Brazilian coffee. Talanta 2018, 176, 59–68. [CrossRef] [PubMed]
33. DeAraújo, T.K.L.; Nobrega, R.O.; Fernandes, D.D.S.; DeAraújo, M.C.U.; Diniz, P.H.G.D.; DaSilva, E.C. Non-destructive au-
thentication of Gourmet ground roasted coffees using NIR spectroscopy and digital images. Food Chem. 2021, 364, 130452.
[CrossRef]
34. Okubo, N.; Kurata, Y. Nondestructive Classification Analysis of Green Coffee Beans by Using Near-Infrared Spectroscopy. Foods
2019, 8, 82. [CrossRef]
35. Zhu, M.; Long, Y.; Chen, Y.; Huang, Y.; Tang, L.; Gan, B.; Yu, Q.; Xie, J. Fast determination of lipid and protein content in green
coffee beans from different origins using NIR spectroscopy and chemometrics. J. Food Compos. Anal. 2021, 102, 104055. [CrossRef]
36. Giraudo, A.; Grassi, S.; Savorani, F.; Gavoci, G.; Casiraghi, E.; Geobaldo, F. Determination of the geographical origin of green
coffee beans using NIR spectroscopy and multivariate data analysis. Food Control 2019, 99, 137–145. [CrossRef]
37. Marquetti, I.; Link, J.V.; Lemes, A.L.G.; Scholz, M.B.S.; Valderrama, P.; Bona, E. Partial least square with discriminant analysis and
near infrared spectroscopy for evaluation of geographic and genotypic origin of arabica coffee. Comput. Electron. Agric. 2016, 121,
313–319. [CrossRef]
38. Santos, J.R.; Lopo, M.; Rangel, A.O.S.S.; Lopes, J.A. Exploiting near infrared spectroscopy as an analytical tool for on-line
monitoring of acidity during coffee roasting. Food Control 2016, 60, 408–415. [CrossRef]
39. Ribeiro, J.S.; Salva, T.J.G.; Silvarolla, M.B. Prediction of a wide range of compounds concentration in raw coffee beans using NIRS,
PLS and variable selection. Food Control 2021, 125, 107967. [CrossRef]
40. Ribeiro, J.S.; Ferreira, M.M.C.; Salva, T.J.G. Chemometric models for the quantitative descriptive sensory analysis of arabica coffee
beverages using near infrared spectroscopy. Talanta 2011, 83, 1352–1358. [CrossRef]
41. Craig, A.P.; Franca, A.S.; Oliveira, S.L.; Irudayaraj, J.; Ileleji, K. Fourier transform infrared spectroscopy and near infrared
spectroscopy for the quantification of defects in roasted coffees. Talanta 2015, 134, 379–386. [CrossRef]
42. Scholz, M.B.S.; Kitzberger, C.S.G.; Pereira, L.F.P.; Davrieux, F.; Pot, D.; Charmetantd, P.; Leroy, T. Application of near infrared
spectroscopy for green coffee biochemical phenotyping. J. Near Infrared Spectrosc. 2014, 22, 411–421. [CrossRef]
43. Alessandrini, L.; Romani, S.; Pinnavaia, G.; Rosa, M.D. Near infrared spectroscopy: An analytical tool to predict coffee roasting
degree. Anal. Chim. Acta 2008, 625, 95–102. [CrossRef] [PubMed]
44. Abreu, M.B.; Marcheafave, G.G.; Bruns, R.E.; Scarminio, I.S.; Zeraik, M.L. Spectroscopic and Chromatographic Fingerprints for
Discrimination of Specialty and Traditional Coffees by Integrated Chemometric Methods. Food Anal. Methods 2020, 13, 2204–2212.
[CrossRef]
45. Craig, A.P.; Botelho, B.G.; Oliveira, L.S.; Franca, A.S. Mid infrared spectroscopy and chemometrics as tools for the classification of
roasted coffees by cup quality. Food Chem. 2018, 245, 1052–1061. [CrossRef] [PubMed]
46. Fioresi, D.B.; Pereira, L.L.; Oliveira, E.C.S.; Moreira, T.R.; Ramos, A.C. Mid infrared spectroscopy for comparative analysis of
fermented arabica and robusta coffee. Food Control 2021, 121, 107625. [CrossRef]
47. Flores-Valdez, M.; Meza-Márquez, O.G.; Osorio-Revilla, G.; Gallardo-Velázquez, T. Identification and Quantification of Adulter-
ants in Coffee (Coffea arabica L.) Using FT-MIR Spectroscopy Coupled with Chemometrics. Foods 2020, 9, 851. [CrossRef]
48. Zhang, C.; Wang, C.; Liu, F.; He, Y. Mid-Infrared Spectroscopy for Coffee Variety Identification: Comparison of Pattern Recognition
Methods. J. Spectrosc. 2016, 2016, 7927286. [CrossRef]
49. Craig, A.P.; Franca, A.S.; Oliveira, L.S. Discrimination between Immature and Mature Green Coffees by Attenuated Total
Reflectance and Diffuse Reflectance Fourier Transform Infrared Spectroscopy. J. Food Sci. 2011, 76, 2011. [CrossRef]
50. Link, J.V.; Lemes, A.L.G.; Marquetti, I.; Scholz, M.B.S.; Bona, E. Geographical and genotypic classification of arabica coffee
using Fourier transform infrared spectroscopy and radial-basis function networks. Chemom. Intell. Lab. Syst. 2014, 135, 150–156.
[CrossRef]
51. Dias, R.C.E.; Valderrama, P.; Março, P.H.; Scholz, M.B.S.; Edelmann, M.; Yeretzian, C. Quantitative assessment of specific defects
in roasted ground coffee via infrared-photoacoustic spectroscopy. Food Chem. 2018, 255, 132–138. [CrossRef]
52. Bona, E.; Marquetti, I.; Link, J.V.; Makimori, G.Y.F.; da Costa Arca, V.; Lemes, A.L.G.; Ferreira, J.M.G.; dos Santos Scholz, M.B.;
Valderrama, P.; Poppi, R.J. Support vector machines in tandem with infrared spectroscopy for geographical classification of green
arabica coffee. LWT–Food Sci. Technol. 2017, 76, 330–336. [CrossRef]
53. Rubayiza, A.B.; Meurens, M. Chemical Discrimination of Arabica and Robusta Coffees by Fourier Transform Raman Spectroscopy.
J. Agric. Food Chem. 2005, 53, 4654–4659. [CrossRef]
54. Keidel, A.; Stetten, D.; Rodrigues, C.; Guas, C.; Hildebrandt, P. Discrimination of Green Arabica and Robusta Coffee Beans by
Raman Spectroscopy. J. Agric. Food Chem. 2010, 58, 11187–11192. [CrossRef]
55. Wermelinger, T.; D’Ambrosio, L.; Klopprogge, B.; Yeretzian, C. Quantification of the Robusta Fraction in a Coffee Blend via
Raman Spectroscopy: Proof of Principle. J. Agric. Food Chem. 2011, 59, 9074–9079. [CrossRef]
56. Figueiredo, L.P.; Borém, F.M.; Almeida, M.R.; Oliveira, L.F.C.; Oliveira, L.F.C.; Alves, A.P.C.; Santos, C.M.; Rios, P.A. Raman
spectroscopy for the differentiation of Arabic coffee genotypes. Food Chem. 2019, 288, 262–267. [CrossRef]
57. Sunarharum, W.B.; Williams, D.J.; Smyth, H.E. Complexity of coffee flavor: A compositional and sensory perspective. Food Res.
Int. 2014, 62, 315–325. [CrossRef]
58. Bertone, E.; Venturello, A.; Giraudo, A.; Pellegrino, G.; Geobaldo, F.G. Simultaneous determination by NIR spectroscopy of the
roasting degree and Arabica/Robusta ratio in roasted and ground coffee. Food Control 2016, 59, 683–689. [CrossRef]
Processes 2022, 10, 71 24 of 25

59. Pires, F.C.; Pereira, R.G.F.A.; Baqueta, M.R.; Valderrama, P.; Rocha, R.A. Near-infrared spectroscopy and multivariate calibration
as an alternative to the Agtron to predict roasting degrees in coffee beans and ground coffees. Food Chem. 2021, 365, 130471.
[CrossRef] [PubMed]
60. Catelani, T.A.; Santos, J.R.; Páscoa, R.N.M.J.; Pezza, L.; Pezza, H.R.; Lopes, J.A. Real-time monitoring of a coffee roasting process
with near infrared spectroscopy using multivariate statistical analysis: A feasibility study. Talanta 2018, 179, 292–299. [CrossRef]
61. Ebrahimi-Najafabadi, H.; Leardi, R.; Oliveri, P.; Casolino, M.C.; Jalali-Heravi, M.; Lanteri, S. Detection of addition of barley to
coffee using near infrared spectroscopy and chemometric techniques. Talanta 2012, 99, 175–179. [CrossRef] [PubMed]
62. Winkler-Moser, J.K.; Singh, M.; Rennick, K.A.; Bakota, E.L.; Jham, G.; Liu, S.X.; Vaughn, S.F. Detection of Corn Adulteration in
Brazilian Coffee (Coffea arabica) by Tocopherol Profiling and Near-Infrared (NIR) Spectroscopy. J. Agric. Food Chem. 2015, 63,
10662–10668. [CrossRef]
63. Brondi, A.M.; Torres, C.; Garcia, J.S.; Trevisan, M. Differential scanning calorimetry and infrared spectroscopy combined with
chemometric analysis to the determination of coffee adulteration by corn. J. Braz. Chem. Soc. 2017, 28, 1308–1314. [CrossRef]
64. Santos, J.R.; Sarraguça, M.C.; Rangel, A.O.S.S.; Lopes, J.A. Evaluation of green coffee beans quality using near infrared spec-
troscopy: A quantitative approach. Food Chem. 2012, 135, 1828–1835. [CrossRef]
65. Chen, S.; Chang, C.; Ou, C.; Lien, C. Detection of Insect Damage in Green Coffee Beans Using VIS-NIR Hyperspectral Imaging.
Remote Sens. 2020, 12, 2348. [CrossRef]
66. Craig, A.P.; Franca, A.S.; Oliveira, L.S. Evaluation of the potential of FTIR and chemometrics for separation between defective
and non-defective coffees. Food Chem. 2012, 132, 1368–1374. [CrossRef]
67. Baqueta, M.R.; Coqueiro, A.; Valderrama, P. Brazilian coffee blends: A simple and fast method by near-infrared spectroscopy for
the determination of the sensory attributes elicited in professional coffee cupping. J. Food Sci. 2019, 84, 1247–1255. [CrossRef]
68. Chang, Y.T.; Hsueh, M.C.; Hung, S.P.; Lu, J.M.; Peng, J.H.; Chen, S.F. Prediction of specialty coffee flavors based on near-infrared
spectra using machine and deep-learning methods. J. Sci. Food Agric. 2021, 101, 4705–4714. [CrossRef]
69. Tolessa, K.; Rademaker, M.; Baets, B.D.; Boeckx, P. Prediction of specialty coffee cup quality based on near infrared spectra of
green coffee beans. Talanta 2016, 150, 367–374. [CrossRef] [PubMed]
70. Arboleda, E.R. Discrimination of civet coffee using near infrared spectroscopy and artificial neural network. Int. J. Adv. Comput.
Res. 2018, 8, 324–334. [CrossRef]
71. Belchior, V.; Botelho, B.G.; Oliveira, L.S.; Franca, A.S. Attenuated Total Reflectance Fourier Transform Spectroscopy (ATR-FTIR)
and chemometrics for discrimination of espresso coffees with different sensory characteristics. Food Chem. 2019, 273, 178–185.
[CrossRef]
72. Esteban-Díez, I.; González-Sáiz, J.M.; Saenz-Gonzalez, J.M.; Pizarro, C.I. Coffee varietal differentiation based on near infrared
spectroscopy. Talanta 2007, 71, 221–229. [CrossRef] [PubMed]
73. Medina, J.; Caro Rodríguez, D.; Arana, V.A.; Bernal, A.; Esseiva, P.; Wist, J. Comparison of Attenuated Total Reflectance
MidInfrared, Near Infrared, and 1H-Nuclear Magnetic Resonance Spectroscopies for the Determination of Coffee’s Geographical
Origin. Int. J. Anal. Chem. 2017, 2017, 7210463. [CrossRef]
74. Luna, A.S.; DaSilva, A.P.; Alves, E.A.; Rocha, R.B.; Lima, I.C.A.; DeGois, J.S. Evaluation of chemometric methodologies for the
classification of Coffea canephora cultivars via FTNIR spectroscopy and direct sample analysis. Anal. Methods 2017, 9, 4255.
[CrossRef]
75. Wang, N.; Fu, Y.; Lim, L. Feasibility Study on Chemometric Discrimination of Roasted Arabica Coffees by Solvent Extraction and
Fourier Transform Infrared Spectroscopy. J. Agric. Food Chem. 2011, 59, 3220–3226. [CrossRef]
76. Dankowska, A.; Domagała, A.; Kowalewski, W. Quantification of Coffea arabica and Coffea canephora var. robusta concentration
in blends by means of synchronous fluorescence and UV-Vis spectroscopies. Talanta 2017, 172, 215–220. [CrossRef] [PubMed]
77. Escobar, M.M.; Torres, A.E.L.; Rodriguez, N.J.M. Non-Destructive Prediction of Moisture Content of Philippine Coffea arabica and
Coffea liberica Green Beans Using Locally-Developed NIR Spectroscopy Instrument. Mindanao J. Sci. Technol. 2020, 18, 208–223.
78. Kyaw, E.M.; Budiastra, I.W.; Samsudin; Sutrisno. Estimation of moisture content in Liberica coffee by using near infrared
spectroscopy. IOP Conf. Ser. Earth Environ. Sci. 2020, 542, 012013. [CrossRef]
79. Macedo, L.L.; Araújo, C.D.S.; Vimercati, W.C.; Hein, P.R.G.; Pimenta, C.J.; Saraiva, S.H. Evaluation of chemical properties of intact
green coffee beans using near-infrared spectroscopy. J. Sci. Food Agric. 2020, 101, 3500–3507. [CrossRef]
80. Liang, N.; Lu, X.; Hu, Y.; Kitts, D.D. Application of Attenuated Total Reflectance-Fourier Transformed Infrared (ATR-FTIR)
Spectroscopy To Determine the Chlorogenic Acid Isomer Profile and Antioxidant Capacity of Coffee Beans. J. Agric. Food Chem.
2016, 64, 681–689. [CrossRef]
81. Weldegebreal, B.; Redi-Abshiro, M.; Chandravanshi, B.S. Development of new analytical methods for the determination of
caffeine content in aqueous solution of green coffee beans. Chem. Cent. J. 2017, 11, 126. [CrossRef]
82. Yisak, H.; Redi-Abshiro, M.; Chandravanshi, B.S. New fluorescence spectroscopic method for the simultaneous determination of
alkaloids in aqueous extract of green coffee beans. Chem. Cent. J. 2018, 12, 59. [CrossRef] [PubMed]
83. Bressanello, D.; Liberto, E.; Cordero, C.; Rubiolo, P.; Pellegrino, G.; Ruosi, M.R.; Bicchi, C. Coffee aroma: Chemometric comparison
of the chemical information provided by three different samplings combined with GC–MS to describe the sensory properties in
cup. Food Chem. 2017, 214, 218–226. [CrossRef]
84. Alcantara, G.M.R.N.; Dresch, D.; Melchert, W.R. Use of non-volatile compounds for the classification of specialty and traditional
Brazilian coffees using principal component analysis. Food Chem. 2020, 360, 130088. [CrossRef] [PubMed]
Processes 2022, 10, 71 25 of 25

85. Toci, A.T.; Farah, A.; Pezza, H.R.; Pezza, L. Coffee Adulteration: More than Two Decades of Research. Crit. Rev. Anal. Chem. 2016,
46, 83–92. [CrossRef] [PubMed]
86. Forchetti, D.A.P.; Poppi, R.J. Detection and Quantification of Adulterants in Roasted and Ground Coffee by NIR Hyperspectral
Imaging and Multivariate Curve Resolution. Food Anal. Methods 2020, 13, 44–49. [CrossRef]
87. Franca, A.S.; Oliveira, L.S.; Mendonc, J.C.F.; Silva, X.A. Physical and chemical attributes of defective crude and roasted coffee
beans. Food Chem. 2005, 90, 84–89. [CrossRef]

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy