ChapterPDF Available

Pears

Authors:

Abstract and Figures

The most important commercial pear species grown are Pyrus communisand P. pyrifolia, although there are significant acreages of several other species. Climatic adaptation is a concern of all pear breeders, as well as fruit quality, season extension, compatibility with major pollenizer cultivars and disease resistance. Increasing fire blight resistance is an important goal in the eastern and southern parts of North America, and many regions of Europe. In the breeding of pear rootstocks, the common goal is to develop rootstocks that induce size control and precocity in the scion cultivar. Resistance to fire blight is quantitatively inherited in an additive fashion, with a few major genes playing an important role. Genes have also been identified for semidwarf or compact cultivars, short-internode dwarfs and short internode compact pears. Varying amounts are known about the genetics of fruit development and quality. Molecular studies have been conducted on the expression patterns of genes during fruit ripening and storage. Several genetic maps have been developed of pear and DNA markers have been linked to a number of resistance genes including black spot disease, scab and fire blight resistance. Transformation strategies have been employed to generate herbicide and disease resistant plants.
Content may be subject to copyright.
1
Chapter 10
Pears
J. F. Hancock 1, G. A. Lobos 2
1 Department of Horticulture, 342C Plant and Soil Sciences Building, Michigan State
University, East Lansing, Michigan 48824, U.S.A. 2 Laboratorio de Ecofisiologia Vegetal,
Escuela de Agronomia, Universidad de Talca, Chile.
E-mail: hancock@msu.edu
Abstract
The most important commercial pear species grown are Pyrus communis and P. pyrifolia, although
there are significant acreages of several other species. Climatic adaptation is a concern of all pear
breeders, as well as fruit quality, season extension, compatibility with major pollenizer cultivars and
disease resistance. Increasing fire blight resistance is an important goal in the eastern and southern
parts of North America, and many regions of Europe. In the breeding of pear rootstocks, the common
goal is to develop rootstocks that induce size control and precocity in the scion cultivar. Resistance to
fire blight is quantitatively inherited in an additive fashion, with a few major genes playing an important
role. Genes have also been identified for semidwarf or compact cultivars, short-internode dwarfs and
short internode compact pears. Varying amounts are known about the genetics of fruit development
and quality. Molecular studies have been conducted on the expression patterns of genes during fruit
ripening and storage. Several genetic maps have been developed of pear and DNA markers have
been linked to a number of resistance genes including black spot disease, scab and fire blight
resistance. Transformation strategies have been employed to generate herbicide and disease
resistant plants.
10.1 Introduction
Pears are grown in all temperate regions of the world. Culivars of the European pear, Pyrus
communis, predominate in Europe, North America, South America, Africa and Australia, and
the sand or Japanese pear, P. pyrifolia, is the main cultivated species in southern and central
China, Japan and Southeast Asia. Other pears grown widely in Asia include P. ussuriensis
(Ussuri pear) and hybrids of P. pyrifolia and P. ussuriensis. Interest in Asian pears, primarily
Japanese cultivars, continues to increase in Western Europe, North America, New Zealand
and Australia, but the European pear has made little impact in Asia, except in Northern Japan,
where most Asian pears do not have sufficient winter hardiness. Pears are eaten fresh,
cooked, dried or made into a fermented cider-like beverage called “perry”. They are also
processed as canned halves, diced pieces for fruit cocktail and as puree for baby food.
The world production of pears is second only to apples among the deciduous tree fruits.
About 19.2 million metric tons of pears were produced in 2005 (FAOSTAT, 2007), up over 2
million tons from 2004. Asia produced the most pears (13.5 million t), followed by Europe
(3.2 million t), South America (0.8 million t), North America (0.8 million t), Africa (0.7
million t), and Oceania (0.2 million t). In Asia, China was the largest producer with 60% of
the world volume, followed by Japan (2.1%) and Republic of Korea (2.0%). The major
producers in Europe were Italy (4.9%), Spain (3.4%) and France (1.2%). In South America,
the highest producers were Argentina (2.7%) and Chile (1.1%). In North America, the
2
leading producer by far was the U.S.A. (3.9%), where the main state is Washington with 50
percent of the countries volume (California and Oregon complete the other half). In Africa,
the major producer was the Republic of South Africa (1.8%). Production in Oceania was
highest in Australia (0.8%).
Pyrus is genetically quite diverse, with considerable variability in morphology and
physiological adaptations (Figure 10.1) (Knight 1963, Westwood 1982, Lombard and
Westwood 1987, Bell et al. 1996). European and Asian pear breeders have utilized this
variability to develop high quality cultivars with large size and attractive appearance that are
well adapted to local conditions. The European pears are distinguished by their juiciness,
delicate flavor and aroma, while the Oriental (Asian or nashi) pears are known for their
crispness and sweet flavor. North American breeders have had to focus more on disease
resistance and cold hardiness than the Europeans, although the spread of the bacterial disease
fire blight [Erwinia amylovora (Burrill) Winslow et al.] throughout Europe is forcing the
breeders there to concentrate more on disease resistance (Bell et al. 1996).
Figure 10.1. Diversity in fruit of Pyrus species (Picture by Joseph Postman, USDA-ARS
National Clonal Repository in Corvallis, Oregon.
10.2 Evolutionary biology and germplasm resources
The genus Pyrus is in the subfamily Pomoideae of the Rosaceae. All species of Pyrus are
diploid and interfertile (Westwood and Bjornstad 1971, Bell and Hough 1986). The genus
contains 24 primary species (Table 10.1), up to six natural interspecific hybrids and at least
three artificial hybrids (Bell et al. 1996). The species of Pyrus are located in Europe,
temperate Asia and the mountains of North Africa. The species boundaries are blurred in
many instances, resulting in different species designations by some authorities.
Pomoideae are unique in the Rosaceae by having a basic chromosome number of 17
compared to 7 to 9 for the other subfamilies. The origin of the subfamily may have occurred
when two primitive forms of Rosaceae successfully hybridized, one having a basic
chromosome number of 8 and the other 9 (Sax 1931, Zielinski and Thompson 1967). These
could have been members of the Prunoideae and Spiraeoideae. All the species of Pyrus have
a chromosome number of 2x = 34, except for a few higher polyploid cultivars of P.
3
communis. The genus Pyrus probably arose during the Tertiary period in the mountains of
western China and the early taxa likely dispersed east and west through the mountain chains.
Speciation may have been associated with geographical isolation of populations in the
mountain ranges (Rubtsov 1944, Zeven and Zhukovsky 1975).
Table 10.1 Pyrus species of the world (adapted from Bell et al. 1996).
Scientific name Synonyms Distribution
European
P. communis L. P. asiae-mediae Popov
P. balansae Decne.
P. boissieriana Buhse
P. caucasica Fed
P. elata Rubtzov
P. medvedevii Rubtzov
West to southeast Europe,
Turkey
P. korshinskyi Litv. P. pyraster (L.) Burgsd. South central Asia, Afghanistan
P. nivalis Jacq. West, Central and Southern
Europe
P.cordata Desv. S.W. England, W. France,
Spain, Portugal
P. ×salviifolia DC Europe, Crimea
Circum-Mediterranean
P. amygdaliformis Vill P. sinaica Dum. Cours. Mediterranean Europe, Asia
minor
P. complexa Rubtzov Caucasus
P. elaeagrifolia Pall. P. kotschyana Boiss. ex Decne. S.E. Europe, Russia, Turkey
P. syriaca Boiss. Tunisia
P. longipes Coss. & Dur. Algeria
P. gharbiana Trab. Morocco, W. Algeria
P. mamorensis Trab. Morocco
Mid-Asian
P. glabra Boiss. Iran
P. salicifolia Pall. N.W. Iran, N.E. Turkey, South
Russia
P. regelii Rehd. P. heterophylla Regal & Schmalh. South Central Asia
P. pashia Buch.-Ham. ex D. Don. P. kumaoni Decne.
P. variolosa Wall. ex G. Don.
P. wilhelmii C.K. Schneid.
Pakistan, India, Nepal
East Asian
P. ×bretschneideri Rehd. Northern China
P. ×phaeocarpa Rehd. Northern China
P. pyrifolia (Burm.)Nak P. serotina Rehd. China, Japan, Korea
P. ×serrulata Rehd. Central China
P. pseudopashia T.T. Yu P. kansuensis Batalin N.W. China
P. ussuriensis Maxim P. lindleyi Rehd.
P. ovoidea Rehd.
P. sinensis Lindley
Siberia, Manchuria, N. China,
Korea
P. calleryana Decne. Central & S. China, Vietnam
P. betulifolia Bunge Central & N. China, S.
Manchuria
P. fauriei C.K. Scheid. Korea
P. hondoensis Kik. & Nak Japan
P. dimorphophylla Mak. Japan
P. kawakamii Hayata P. koehnei C.K. Schneid. Taiwan, S.E. China
4
A number of pear species are grown commercially (Bell et al. 1996). As mentioned
previously, Pyrus communis (common pear) is the most important pear in Europe, North
America, South America, Africa, and Australia. Pyrus nivalis, the snow pear, is grown
locally in Europe to make perry and hybrids of P. communis and P. pyrifolia are grown in
North America for processing. The sand pear, P. pyrifolia, is the primary pear species
cultivated in southern and central China and in Japan, with hectarage of the Ussuri pear, P.
ussuriensis, hybrids of P. pyrifolia and P. ussuriensis, and the Chinese white pear (P. ×
bretschneideri) also being found in northern China and Japan. Pyrus pashia (Pashia pear) is
grown in southern China and northern India. Several species are used for rootstocks including
P.betulifolia, P. calleryana, P. pyrifolia, P. ussuriensis, and P. communis in Europe, North
America and eastern Asia, and P. pyraster, P. amygdaliformis and P. elaeagrifolia in Asia
Minor and central Asia. The small fruited species, P. calleryana, P. fauriei, P. betulifolia, P.
salicifolia, and P. kawakamii are grown as ornamentals.
The fruit of the Asian pears are notable in that they can be eaten right after harvest, unlike the
European pears. They are sweet and juicy and tend to have crisp-textured flesh, but do not
have the smooth, buttery texture of European pears, tend to have less aroma, and some have
abundant stone cells (Bell et al. 1996). The Chinese or Japanese sand pear (P. pyrifolia) is
well adapted to warm climates and is very widespread with hundreds of cultivars grown. The
fruit vary in size from very small to very large and they are sweet, juicy and gritty, with a
russeted skin. Pyrus pashia is well adapted to very hot, humid climates, but has only
mediocre fruit quality compared to the other species.
The Ussurian pear, P. ussuriensis, is the most cold hardy Asian pear and is grown in Northern
Japan where P. pyrifolia is susceptible to winter injury. It has small, globose medium-sized
fruit with persistent calyxes. They are bland flavored but not too gritty and some have
melting flesh (Shen 1980). The Chinese white pear, P. ×bretschneideri is the second most
hardy Asiatic type. Their fruits are of medium size, range in shape from pyriform to obovate
and have the most pleasing texture and flavor of the Oriental pears.
It is likely that interspecific hybridization played an important role in pear domestication. A
number of species have been implicated as being in the background P. communis including
wild populations of P. communis var. pyraster, P. caucasica and P. nivalis (Challice and
Westwood 1973). Rubtsov (1944) felt that modern cultivars of P. communis had
characteristics derived from at least three species, P. elaeagrifolia, P. salicifolia and P.
syriaca. Cultivars grown in northern China may belong to a hybrid complex involving P.
ussuriensis and P. pyrifolia (Bell et al. 1996). The origin of P. ×bretschneideri likely
involved the hybridization of P. betulnefolia and P. ussuriensis or P. pyrifolia (Kikuchi
1946). The most recent molecular data suggests that P. ×bretschneideri is closely related to
P. pyrifolia and P. ussuriensis, and is likely a variety or subspecies of P. pyrifolia
(Yamamoto et al. 2002a, Bao et al. 2007).
Vavilov (1951) identified three centers of diversity for cultivated pears: 1) A Chinese center,
where P. pyrifolia and P. ussuriensis are found. The primitive species, P. calleryana, is
located in this center (Challice and Westwood 1973, Pu et al. 1986). 2) A central Asiatic
center (northwest India, Afghanistan, Tadjikistan and Uzbekistan), where P. communis and
hybrids of P. communis and P. ×bretschneideri are found (Yu and Zhang 1979). 3) A Near
Eastern center (Caucasus Mountains and Asia Minor), where P. communis is also grown. It is
in the Near Eastern center that P. communis may have been first domesticated.
5
The USDA-ARS National Plant Germplasm System (NPGS) clonal repository at Corvallis,
maintains an extensive and diverse pear germplasm collection that includes historic varieties,
landrace varieties, and wild species (more than 2,200 different individuals in total). This
collection includes genotypes with desirable traits important for crop improvement, such as
resistance to fire blight, pear psylla, Fabraea leaf spot, and pear scab.
10.3 History of improvement
Originating in the Caucasus region (Southeastern Europe between Black and Caspian seas),
pears have been cultivated for at least 3,000 years in Asia (Kikuchi 1946). Indo-European
tribes spread the pear as they migrated into European and Northern India. The first mention
of pears in the written record was made by Homer in about 1000 B.C., when he referred to
them as one of the “gifts of the Gods” (Hedrick et al. 1921). According to Theophrastus (371-
286 B.C.), pear culture was common in ancient Greece, where cultivars were propagated by
grafting and cuttings. The Roman, Cato (235-150 B.C.), described pear cultural methods that
are very similar to the techniques practiced today. By the time of Pliny the Elder (23-79 BC),
at least 35 cultivars of pear were being grown in Rome, while only three types of apples were
noted. The range in fruit characters of these ancient cultivars were similar to those grown
today. By the end of the Sung Dynasty (China, AD 1279), over one hundred varieties of pear
existed.
In the Middle Ages, pears were grown widely in central and western Europe (Bell et al.
1996). Cordus (1515-1544) described pear cultivars in Germany that had all the fruit
characters possessed by modern cultivars, with the exception of buttery texture. France was
the leading pear-producing country in the sixteenth and seventeenth centuries, and were most
active in the development of new cultivars. In 1628, the amateur fruit collector Le Lectier had
254 pear cultivars in his garden and by the early 1800s there were over 900 cultivars of pears
growing in France. Most of these were crisp-fleshed.
In the eighteenth century, Belgium became the center of pear culture and improvement,
primarily through the efforts of Nicolas Hardenpont and Jean Baptiste Van Mons (Bell et al.
1996). The Belgian breeders developed the first cultivars that had a melting buttery flesh, and
some of their cultivars are still important today, including 'Beurre Bosc', 'Beurre d’Anjou' and
'Winter Nelis'.
The pear was not native to England, but commercial culture probably arose there by 1200. It
is not known when the pear was introduced into England, but it is likely to have been before
the Roman conquest (Bell et al. 1996). The English became expert pear breeders in the 1800s,
evidenced by an 1826 catalog of the Royal Society of London listing 622 cultivars. The most
important cultivar now grown in the world, 'Williams Bon Chretien' ('Bartlett'), was identified
around 1796 (Figure 10.2) (Hooker 1818). Another 19th century selection named
'Conference', is also still planted.
Only two species were bred in Europe, Pyrus communis (the common European pear), and P.
nivalis (the perry pear). Most of the improved types were derived from open-pollinated
seedlings of existing cultivars, although Knight in England was using controlled
hybridization around 1800 (Bell et al. 1996). The father of modern genetics, Gregor Mendel,
was also hybridizing pears at Brno in the late nineteenth century to develop late-ripening
cultivars with superior flesh quality (Vavra and Orel 1971).
6
Figure 10.2. The 'Bartlett' pear (watercolor published in The Fruits of New York, U.P.
Hedrick, 1921).
Domestication of the Chinese pear species began about 3300 years ago (Kikuchi 1946) and
commercial orchards have existed in China for more than 2000 years (Pieniazek 1966).
However, pear breeding did not begin in China until 1956. Pears were cultivated as early as
the eighth century in Japan (Kajiura 1966), with large plantings not appearing until 1868.
Kikuchi began the first breeding program in Japan in 1915, which is still active (Kanato et al.
1982).
Pears were introduced to North America by the early French and English settlers; the first
reference to pear culture was made in New England in 1629 (Hedrick et al. 1921). Pears were
introduced into the west Coast by the Franciscan monks, and by 1800, pear cultivation
extended from British Columbia to southern California. In the 17th and 18th centuries, only P.
communis was grown in North America and it was not until the early 1800s that P. pyrifolia
arrived at the west coast of the United States via Chinese immigrants. The first hybrid of
these two species, 'Le Conte', appeared in 1846, followed by 'Kieffer' (1873) and 'Garber'
(1880) (Hedrick et al. 1921). The fruit of these hybrids was not of as high quality as the
existing European cultivars, but the trees were more resistant to fire blight. These early
hybrids were likely accidental, but controlled crosses were soon being made in hopes of
combining high fruit quality with blight resistance.
Cold hardy cultivars of P. communis were introduced to North America from Russia in 1879.
These trees had poor fruit quality and were susceptible to fire blight, but they proved to be an
excellent source of cold hardiness (Magness 1937). Additional sources of cold hardiness
came with the introduction of P. ussuriensis into Iowa by Patten around 1867. Chance
hybrids of P. ussuriensis x P. communis proved to be very cold hardy. Reimer (1925)
introduced a number of European and Oriental cultivars to North America in the early 1900s,
and some of the Oriental introductions had superior blight resistance. This stimulated the U.S.
7
Department of Agriculture and several State Agricultural experiment stations to begin
actively breeding for fire blight resistance (Bell et al. 1996).
In general, pear breeding is only a recent activity in South America, Australia, New Zealand
and Africa. However, 'Packham’s Triumph', was introduced commercially in Australia
around 1900 and is still important there and in New Zealand (Lombard and Westwood 1987).
10.4 Current breeding efforts
Climatic adaptation is a concern of all pear breeders. Cold hardiness is of paramount
importance in the more temperate and most northern production regions, while in the warmer
regions the emphasis is on drought and heat tolerance. A goal of most pear breeders is to
generate a series of cultivars that produce a continuous supply of quality fruit throughout the
season. Compatibility with major pollenizer cultivars is important everywhere with an
overlapping bloom period.
Disease resistance is also very important to most pear breeders. Increasing fire blight
resistance is a major goal in the eastern and southern parts of North America, and many
regions of Europe. Resistance is also sought in many areas to pear psylla [Cacopsylla
pyricola (Foerster)], powdery mildew [Podosphaera leucotricha (Ellis & Everh.) E.S.
Salmon], leaf spot and scab (Venturia pirina Aderh.). In Japan, resistance to black spot of
Asian pear (Alternaria alternata (Fr.) Keissler) is particularly important, as well as resistance
to the Asian pear scab (Venturia nashicola Tan. & Yan.) and pear rust (Gymnosporangium
asiaticum Miyabe ex Yamada).
All pear breeders seek high fruit quality, although what constitutes good quality varies by
species and location, as was previously mentioned. A soft, buttery flesh is desired in
European pears along with aromatic flavor and low fiber. A crisp, breaking flesh is preferred
in Oriental pears along with a sweet flavor and lack of grittiness. In European pears, the ideal
fruit is considered large (7 cm long and 6 cm wide) with a pyriform shape. The skin can be a
wide range of colors from golden yellow with or without a red blush, green or greenish-
yellow and even bright red. Russet free skin is generally preferred, although russeting that is
uniform and smooth is acceptable. Fruit are sought that are resistant to bruising, white fleshed
and contain limited quantities of stone cells. In Asian pears, an even larger size is desired (up
to 10 cm wide), along with a regular, round shape (Kanato et al. 1982). The skin can be
yellow or a light green, glossy and uniformly covered with a golden, smooth russet. In both
European and Oriental pears, uniformity of ripening, a long postharvest storage life and a low
susceptibility to physiological disorders are all important.
A trend that is of increasing importance is the ability of a cultivar to be harvested
mechanically both for fresh markets and processing. Pear cultivars adapted to this approach
need to produce fruit of uniform size and maturity that separates easily on shaking, and has a
tough skin that resists bruising (Bell et al. 1996). It is likely that trees adapted to mechanical
harvesting will need to be adapted to trellises or other support systems, and have trunks and
limbs that can withstand the vibrations associated with mechanical shaking.
In the breeding of pear rootstocks, the common goal is to develop rootstocks that induce size
control and precocity in the scion cultivar. They should also be compatible, winter hardy, and
disease resistant. Adaptation to the specific climatic conditions of each production region is
8
also very important. A reduction in root suckering contributes to decreased risk of infections
and herbicide damage. Because seedlings rootstocks are derived from parent cultivars which
are highly heterozygous and self-incompatible, the rootstocks are genetically not identical,
although they are usually uniform in all important characteristics. Japanese and Chinese
breeding programs have focused on reducing the physiological disorder “black end”. In
Europe, a strong emphasis has been placed on improved cold hardiness in Pyrus and
Cydonia, along with adaptation to iron chlorosis in quince (Chevreau and Bell 2006). Several
disease problems of rootstocks receive primary attention in Europe and North America
including fire blight and the pear decline phytoplasma.
Much effort was devoted in the past on improving quince rootstocks, but more energy is now
being devoted to developing dwarfing Pyrus rootstocks. There is less incompatibility
encountered between intrageneric than intergeneric grafts, and some of the Pyrus species are
more winter hardy, disease resistant, drought tolerant and have better anchorage than the
quince rootstocks (Bell et al. 1996). Only a modest amount of root stock breeding is being
conducted in the United States, but active programs exist in England, France, Italy, Sweden,
the Soviet Union and Romania.
Numerous pear breeding programs are found across the world. Some of the largest programs
in North America are those of the U.S. Department of Agriculture at Kearneysville, West
Virginia and Agriculture and Agri-Food Canada at Harrow, Ontario. At least three private
companies are also producing pear cultivars in the U.S.A. Some of the most important
programs in Europe are at the Institute National de la Recherche Agronomique at Angers and
Dax, France, the Instituto Sperimentale per la Frutticoltura in Rome and Forli, Italy, and the
East Malling Research Station in England. Other pear breeding programs are found in
Norway, Sweden and Romania.
In Asia, major programs exist in China at Xingcheng (Liaoning Province), concentrating on
P. ×bretschneideri, and Zhengzhou (Henan Province) and Hangzhou (Zhejiang Province),
focusing on P. pyrifolia (Wang 1990). Breeding is also being done in Korea at the National
Horticultural Research Institution (Kim and Ko 1991 and 1992) and in Taiwan (Hsu and Lin
1987). In Japan, the largest breeding program is at Kikuchi (Kanato et al. 1982).
In the southern hemisphere, only a few breeding programs exist. A major program was
established in 1983 at HortResearch in New Zealand that focuses on fresh fruit for export to
the Northern Hemisphere. In the 1990s, breeding work was also begun in South Africa
(Human 2005) and Brazil (Barbosa et al. 2007).
10.5 Genetics of economically important traits
10.5.1 Pest and disease resistance
Fire blight is the most important disease of pears in North America and Western Europe. In
fact, commercial pear growing has been largely abandoned in the warm humid regions of the
southeastern, southern, and central United States because of the severity of this problem. The
most widely grown cultivars are generally susceptible, but several new varieties have been
released that have improved resistance (Quamme and Spearman 1983, Bell et al. 2002,
Hunter et al. 2002a and b). 'Seckel' and 'Old Home' were the original sources of resistance.
Resistance is quantitatively inherited in an additive fashion, with a few major genes playing
9
an important role (Decourtyne 1967, Quamme et al. 1990). The highest levels of resistance
are found in the Asian species P. calleryana and P. ussuriensis (Table 10.2).
Table 10.2. Useful horticultural traits carried by native pear species (from Bell 1991).
Scientific name Useful characteristics
European
P. communis L. Resistance to fire blight, white spot, pear scab, pear decline, pear psylla;
cold hardy
P. nivalis Jacq. Resistance to pear psylla; cold hardy
P.cordata Desv. Resistance to codling moth; adaptation to warm winters, high pH,
drought
Circum-Mediterranean
P. amygdaliformis Vill Resistance to codling moth; adaptation to warm winters, high pH,
drought and clay soils
P. elaeagrifolia Pall. Resistance to codling moth; cold hardy
P. syriaca Boiss. Adaptation to drought
P. longipes Coss. & Dur. Cold hardy; adaptation to wet and dry soils
P. gharbiana Trab. Adaptation to drought
P. mamorensis Trab. Adaptation to low pH
Mid-Asian
P. glabra Boiss.
P. salicifolia Pall. Adaptation to drought
P. regelii Rehd. Resistance to pear psylla
P. pashia Buch.-Ham. ex D. Don. Adaptation to warm winters, low pH, clay soils
East Asian
P. pyrifolia (Burn.) Nak Resistance to pear scab
P. pseudopashia T.T. Yu Resistance to codling moth; winter hardy
P. ussuriensis Maxim Resistance to fire blight, pear psylla; winter hardy
P. calleryana Decne. Resistance to fire blight, Fabraea leaf spot, pear psylla, codling moth;
adaptation to warm winters, low pH, wet, dry and clay soils
P. betulifolia Bunge Resistance to pear psylla, codling moth; adaptation to warm winters, low
pH, wet, dry and clay soils
P. fauriei C.K. Scheid. Resistance to fire blight, pear psylla, codling moth; adaptation to warm
and cold winters, low pH, wet and clay soils
P. hondoensis Kik. & Nak Resistance to fire blight
P. dimorphophylla Mak. Resistance to powdery mildew, pear psylla, codling moth; adaptation to
low pH and wet soils
P. kawakamii Hayata Resistance to codling moth; adaptation to warm winters
Identifying resistance to fire blight has been problematic due to a number of factors,
including 1) age, vigor, succulence, and kind of tissue infected; 2) temperature and humidity
relationships during the pre- and post infection period; 3) inoculum purity and concentration;
4) method of inoculation, and 5) virulence of isolates (Bell et al. 1996). A useful rating
system for estimating blight resistance in the field was devised by van der Zwet et al. (1970).
Leaf blight caused by the fungus Fabraea maculata Atk. is a widespread problem in pears,
along with Pseudomonas blight (also called bacterial canker and blossom blast) caused by
Pseudomonas syringae pv. syringae van Hall. Leaf blight resistant genotypes are relatively
common in a number of Pyrus species, while resistance to Pseudomonas blight is limited in
European and Asian pears (Bell et al. 1996). Red-skinned mutants of 'Beurre d’Anjou' have
been reported to be more resistant to Pseudomonas blight than green-skinned ones
(Whitesides and Spotts 1991).
10
Table 10.3. Genetics of pest and disease resistance in pear.
Pest or disease Observations and sources
Bacterial
Fire blight
Erwinia amylovora
Resistance in P. communis is rare but exists (van der Zwet and Keil
1979, Quamme and Spearman 1983, Thibault 1987, van der Zwet and
Bell 1990, Hevesi et al. 2004), High levels of resistance are present in P.
calleryana and P. ussuriensis, but not immunity (Hartman 1957),
Resistance is quantitatively inherited, although major genes have been
implicated (Layne et al. 1968, Thompson et al. 1975, Bell et al. 1977),
Additive effects predominate (Quamme et al. 1990)
Fungi
Black spot
Alternaria alternata
Many sources of resistance exist (Hiroe et al. 1958, Kanato et al. 1982),
Susceptibility is controlled by a single dominant gene (Kozaki 1973)
Leaf blight
Pseudomonas syringae
Resistance has been identified in most Pyrus species (Bell et al. 1996)
Pear scab
Venturia pirina
Resistant genotypes have been found in P. communis, P. pyrifolia, P.
ussuriensis, P. nivalis (Kovalev 1963, Westwood 1982, Bell 1991),
Resistance varies regionally due to the existence of multiple fungal
biotypes (Shabi et al. 1973, Vondracek 1982)
Venturia nashicola The major cultivars are not resistant, but many minor cultivars of many
species carry resistance (Kanato et al. 1982)
Pear rust
Gymnosporangium
haraeanum
Resistance has been identified (Kanato et al. 1982)
Powdery mildew
Podosphaera leucotricha
Resistance has been identified (Fisher 1922, Kanato et al. 1982,
Westwood 1982, Serdani et al. 2006)
Pseudomonas blight
Pseudomonas syringae
Most European and Asian pears are susceptible, red-skinned mutants of
'Beurre d’Anjou' are less susceptible than green-skinned ones
(Whitesides and Spotts 1991)
Phytoplasma
Pear decline Pyrus communis is generally tolerant, P. betulifolia is very tolerant,
while P. pyrifolia, P. ussuriensis and P. calleryana are susceptible
(Lombard and Westwood 1987), Resistance is inherited additively
(Westwood 1996)
Insects
Pear psylla
Cacopsylla spp.
Resistance is common in P. betulifolia, P. calleryana, P. fauriei, P.
ussuriensis, P. ×bretschneideri and P. communis x P. ussuriensis hybrids
(Westigard et al. 1970, Quamme 1984, Bell 1991), resistant genotypes
have been identified in P. nivalis and P. communis (Bell and Stuart 1990,
Bell 1992), resistance was positively correlated with large fruit size in
progeny of P. communis x P. ussuriensis (Harris and Lamb 1973)
Pear sawfly
Caliroa cerasi
Variation in resistance exists (Shaw et al. 2004)
Wooly pear aphid
Eriosoma pyricola
Most pear species are resistant or have variable resistance (Westwood
and Westigard 1969), Immunity has been found in two Indian rootstocks
of P. pashia (Khan 1955)
Nematodes
Root-knot
Meloidogyne spp.
Resistant genotypes of P. communis have been reported (Tufts and Day
1934)
Two species of Venturia cause widespread scab diseases in pears: 1) V. pirina Aderh, which
is particularly important in Europe, but is found everywhere P. communis is grown, and 2) V.
11
nashicola Tanaka & Yamamoto which infects Asian pears throughout their range. Resistant
genotypes have been found to V. pirina in P. communis, P. pyrifolia, P. ussuriensis and P.
nivalis (Kovalev 1963, Westwood 1982, Bell 1991) and resistant cultivars have been released
(Fischer and Mildenberger 2000, Bellini et al. 2000). However, resistance can be variable
regionally due to the existence of multiple fungal biotypes (Shabi et al. 1973, Vondracek
1982). Few major Asian pear cultivars are resistant to V. nashicola, although a number of
minor ones carry resistance (Kanato et al. 1982).
Other locally important diseases that effect pears are black spot [Alternaria alternata (Fr.)
Keissler], Monilia fruit rot [Monilinia fructicola (Wint.) Honey], Fabraea leaf spot (Fabraea
maculata Atk.), powdery mildew [Podosphaera leucotricha (Ell. & Ev.) Salm.], Asiatic pear
rust [Gymnosporangium haraeanum Syd.], along with several virus diseases including pear
bud drop, stony pit and ring spot mosaic. Black spot resistant cultivars have been released
(Kanato et al. 1982, Kajiura 1992) and resistance sources have been identified to all of the
other diseases listed above (Bell et al. 1996, Bell and van der Zwet 2005, Serdani et al. 2006).
The phytoplasma induced disease pear decline is a very widespread problem. It is vectored by
the pear psylla, Cacopsylla spp. (Hibino and Schneider 1970), with C. pyricola Förster being
the causative species in North America, and C. pyri L. and C. pyrisuga Förster being the
vectors in Europe. Pear species vary widely in their resistance to pear psylla, with the east
Asiatic species being more resistant than those from Asia Minor and Europe (Bell et al.
1996). Improved resistance to pear psylla has been developed in Romania (Braniste 2002).
A number of insect pests cause problems in pear orchards (Table 10.3). The woolly pear
aphid (Eriosoma pyricola Baker & Davidson) is a common pest of pear trees in the nursery
and young orchards, particularly in Oregon. Resistance to this aphid is widespread in Pyrus,
with at least eight species carrying resistance (Westwood and Westigard 1969). Other insect
pests that periodically cause problems are the pear leaf blister mite (Eriophyes pyri
Pagenstecher), pear sawfly (Caliroa cerasi L.) and the root-knot nematode (Meloidogyne
spp.). There are no reports of resistance to these pests, except pear sawfly (Shaw et al. 2003
and 2004).
10.5.2 Morphological and physiological traits
Improving winter cold hardiness has played a key role in expanding the range of pear
growing into northern regions of North America, Europe, and Asia. The very cold hardy
species P. ussuriensis from northern Russia has proven extremely useful in the development
of cold hardy cultivars for the northern U.S.A. and Canada (Ronald and Temmerson 1982,
Stushnoff and Garley 1982, Luby et al. 1987, Peterson and Waples 1988), as well as Europe
(Ludin 1942, Zavoronkov 1960, Sansavini 1967). These cold hardy hybrid types do not have
as high quality as the cultivars grown in the main pear-growing regions, but they can
withstand winter temperatures as low as – 30 to – 40 °C and have much higher fruit quality
than their P. ussuriensis parents. Inheritance of cold hardiness has not been studied in pear,
but it is likely to be similar to apple where inheritance is quantitative and mostly additive
(Bell et al. 1996).
Resistance of blossoms to spring frost is also an important goal in cold climates. Two
approaches have been taken to solving this problem: 1) breeding for a late blooming periods,
and 2) identifying genotypes with direct resistance of blossoms to frost. Bloom date has been
12
shown to be highly heritable (Anjou 1954) and late flowering does not appear to be strongly
correlated with late ripening (Baldini 1949). Date of leaf break may be used to predict
flowering date (Bell et al. 1996). A large variation in frost tolerance of blossoms has been
identified, but parental and progeny assessment of frost resistance has proven to be difficult
due to yearly microclimate differences, variations in the impact of fruit set reductions
(Perraudin 1955) and cultivar differences in parthenocarpic fruit set (Simovski et al. 1968).
Reducing the chilling requirement of P. communis is a goal of some southern U.S.A.
breeding programs to expand the range of cultivation. A number of interspecific P. communis
x P. pyrifolia hybrids have been released with low chilling requirement by the University of
Florida. Efforts have also been taken to reduce the chilling requirement of Asian pears at
numerous locations in S. E. Asia, China and India. Detailed information about the chilling
hour requirements of pear varieties can be found at Spiegel-Roy and Alston (1979) and
Ghariani and Stebbins (1994). To our knowledge, no studies have been conducted on the
genetics of chilling requirement in pear.
Drought resistance of scions and rootstocks can be critical when pears are grown in arid to
semiarid conditions. Pruss and Eremeev (1969) found considerable variability in the drought
tolerance of 99 pear cultivars of Russian, European, Australian, and American origin. Pears
grafted onto P. salicifolia performed much better on in dry soils than those on P. communis
rootstocks and had higher tolerance to extreme temperature changes and alkaline soils. They
were also more resistant to pear scab and woolly aphid (Kuznetzov 1941).
Dwarf fruit trees are now favored to standard types, as they are easier to manage and yield
more per area due to more efficient light interception (Tukey 1964). However, there are no
completely dwarf pear cultivars, and potential sources of small size are much harder to find in
pears than in apples. Quince rootstocks are sometimes used to control pear plant size, but the
adaptability of these rootstocks is too limited for widespread use (Bell et al. 1996).
There are several possible approaches to reducing pear tree size (Table 10.4). Semidwarf or
compact cultivars of P. communis have been identified (Tuz 1972) and when they are crossed
with standard types, there appears to be dominance for the compact form. Decourtye (1967)
identified a short-internode dwarf, 'La Nain Vert', that is under single dominant gene control.
This gene has been used in U.S.A. and Italian breeding programs, but the dwarf parent has
poor fruit quality and a very slow growth rate. Jingxian et al. (1988) identified short internode
compact pears from seedlings of 'Jin-xiang' that are also under single gene control. Few
genetically dwarfed scion cultivars have been released to date (Bellini et al. 2000), but active
efforts continue in Italy and France to reduce stature (Rivalta et al. 2002, Chevreau and Bell
2006).
Precocity is very important in pears, both to the breeder who is anxious to evaluate the fruit
and the producer who wants a return on his planting investment as soon as possible. The
length of the juvenile period is heritable and under additive genetic control (Zielinski 1963,
Visser 1967 and 1976, Zimmerman 1976, Bell and Zimmerman 1990). Li et al. (1981) found
that the amount of linear growth to the first flower (“juvenile span”) was positively correlated
with the length of the juvenile period and was heritable. Bell and Zimmerman (1990) found
that in a population of inter-specific hybrid origin, the juvenile period was more dependent on
the individual genotype than species pedigree. Seedlings of P. pyrifolia are more precocious
than P. ×bretschneideri and P. communis (Zhejiang Agricultural University 1978).
13
Table 10.4. Genetics of adaptation, productivity, plant habit and fruit quality in pear.
Attribute Observations and sources
Adaptation
Cold hardiness Quantitatively inherited, largely additive; P. ussurienis is extremely hardy (Bell et al.
1996)
Season of flowering Quantitatively inherited (Anjou 1954, Bell et al. 1996); late flowering only weakly
correlated with late ripening (Baldini 1949)
Harvest date Quantitatively inherited with high heritability observed in P. communis (Thibault et
al. 1988) and moderate heritability in P. pyrifolia (Zhejiang Agricultural University
1977), although some studies have shown parents to be poor predictors of progeny
performance (Crane and Lewis 1949)
Productivity
Juvenile phase length Quantitatively inherited under additive control (Zielinski et al. 1963, Visser 1967 and
1976, Zimmerman 1976, Bell and Zimmerman 1990)
Incompatibility Numerous S-alleles exist, but there are compatible combinations (Sanzol and Herrero
2002, Kim et al. 2004, Moriya et al. 2007, Wu et al. 2007)
Plant habit
Dwarfing Several sources of dwarfing exist that are regulated by single genes (Decourtye 1967,
Jingxian et al. 1988)
Fruit quality
Firmness Quantitatively inherited and highly heritable (Machida and Kozaki 1976, Kajima and
Sato 1990)
Flavor Quantitatively inherited but with low heritability in P. communis (Bell and Janick
1990); poor flavor of P. ussuriensis is inherited quantitatively in interspecific crosses
with some dominance (Lantz 1929); Aroma of P. ussuriensis is highly heritable in
interspecific crosses (Pu et al. 1963)
Juiciness Regulated by a limited number of dominant genes (Zielinski et al. 1965)
Keeping quality Quantitatively inherited (Zielinski et al. 1965)
Russeting Quantitatively controlled in P communis with high heritability (Zielinski et al. 1965,
Crane and Lewis 1949, Bell and Janick 1990); Two genes regulate russetting in P.
pyrifolia, R and I (Kikuchi 1930)
Skin color Major genes determine – yellow dominant to green, blushed recessive to non-blushed
(Zielinski et al. 1965); Deep red regulated by dominant gene C (Zielinski et al. 1963,
Brown 1966); White flesh dominant to colored (Zielinski et al. 1965)
Shape Quantitatively inherited with round and obovate dominant to pyriform and turbinate
(Crane and Lewis 1949, Zielinski et al. 1965, Wang and Wei 1987, White and
Alspach 1996); Moderate to high heritability for ratio of length to diameter (Shin et
al. 1983, White et al. 2000)
Size Quantitatively inherited (Crane and Lewis 1949, Zielinski et al. 1965, Shen et al.
1979, Wang and Wei 1987); Low heritability in some P. pyrifolia populations
(Machida and Kozaki 1976, Shin et al. 1983)
Sugar content Heritability of soluble solids in P. pyrifolia is low (Shin et al. 1983) to moderate
(Machida and Kozaki 1976) depending on population
Texture Quantitatively inherited with relatively high heritability in P. communis (Machida
and Kozaki 1976, Kajiura and Sato 1990); Presence of stone cells dominant to
stoneless in most crosses of Pyrus (Golisz et al. 1971, Pu et al. 1963, Zielinski et al.
1965); Grit content regulated by at least four loci, acting independently and
additively (Thompson et al. 1974)
A number of characters may be useful in the early identification of precocity. Bell et al.
(1996) suggested that juvenile seedlings can be identified by the presence of thorns, irregular
leaf margins and wide-angled branches. Seedlings with short juvenile periods lose these
features quickly. Vigor, measured as stem diameter, is not an accurate predictor of the length
of the juvenile period (Zimmerman 1977, Shen et al. 1982).
14
While there is a need for pear cultivars that ripen all across the season, early ones are
especially important for the fresh market. In P. communis, heritability for date of ripening
was calculated to be 0.49 (Thibault et al. 1988), while in P. pyrifolia a value of 0.33 was
estimated (Zhejiang Agricultural University 1977). Bell et al. (1996) suggested that to breed
most effectively for earliness or lateness, both parents should be either early or late.
10.5.3 Fruit quality
High fruit quality is the prime objective of all fruit breeding programs and encompasses a
wide array of attributes including flavor, texture, appearance, juiciness, postharvest storage
life and incidence of physiological disorders. The most important attribute is eating quality,
which is largely dependent on flavor and texture. Most programs focus on fresh fruit quality,
but some are also interested in the quality of canned or pureed fruit. Considerable attention is
commonly given to core breakdown, bitter pit and superficial scald.
High sugar content is likely the most important factor determining a flavorful pear. The level
of acidity is secondary in importance, as high and low acid pears have been characterized as
good as long as they have high sugar (Visser et al. 1968). Extreme bitterness and astringency
are generally not desirable in dessert pears, although they are important contributors to the
bitter-sharp flavor enjoyed in perry (Luckwill and Pollard 1963). The heritability of flavor in
P. communis has been found to be relatively low (h2 = 0.21) (Bell and Janick 1990).
Subjective perceptions of sweetness and acidity, as well as physical measurements of soluble
solids and acidity, are inherited as independent, quantitative traits (Table 10.4). Estimates of
the heritability of soluble solids in P. pyrifolia cultivars have ranged from quite low (Shin et
al. 1983) to moderate (Machida and Kozaki 1976). Zielinski et al. (1965) found that juiciness,
depending on the parent, was regulated by either a single gene or several, with juiciness being
dominant to dryness. The inheritance of flavor in hybrid populations of P. ussuriensis is
quantitative with poor flavor showing some dominance (Lantz 1929).
It is important that pears can be held in cold storage for long periods of time without
developing internal breakdown, so that they can be sold during the winter when prices tend to
be the highest. Few inheritance studies have been conducted on this characteristic, although
segregation patterns suggest that long term keeping quality is quantitatively inherited
(Zielinski et al. 1965).
A large number of aromatic compounds are produced by pears that make important
contributions to fruit flavor. In 'Bartlett', 77 volatile compounds have been identified with
varying impacts on flavor (Jennings and Tressel 1974). The esters of trans-2, cis-4-
decadienoic acid are highly correlated with the intensity of 'Bartlett' aroma (Jennings et al.
1964), although there are cultivars with high levels of these compounds that do not have that
distinctive aroma (Bell et al. 1996). Seventeen volatile compounds, including toluene, ethyl
butryate, hexanal, (E) 2- hexanal, ethyl hexanoate, a-farnesene, were found to be the most
common constituents of five Asian pear cultivars (Horvat et al. 1992). 'Ya Li' of P. ×
bretschneideri and 'Shinko', a putative interspecific hybrid, lacked five compounds found in
three other P. pyrifolia cultivars ('Chojuro', 'Hosui', and 'Kosui'). The genetics of these
compounds are unknown, although Pu et al. (1963) found that some P. ussuriensis cultivars
had an aromatic quality that was dominant to the sweet, bland flavor of P. ×bretschneideri
and P. pyrifolia cultivars.
15
The desired flesh texture of pears varies from region to region, in part because of species
differences. In Western Europe and North America the favored type is the soft, buttery one
typical of P. communis, while in China and Japan, the preferred texture is the crisp, breaking
one of P. pyrifolia, P. ×bretschneideri, and P. ussuriensis. A minimum of grittiness (stone
cells) is appreciated in all types of pears, although more is tolerated in the Asian than
European ones. Bell and Janick (1990) found heritabilities of 0.30 and 0.45 for texture and
grit scores in P. communis. Flesh firmness has been shown to be highly heritable in
populations of P. pyrifolia (Machida and Kozaki 1976, Kajiura and Sato 1990). Stoneless
flesh was found to be dominant to stony in certain interspecific hybrids of wild Pyrus species
(Westwood and Bjornstad 1971); however, the presence of stone cells was dominant in
several studies of cultivars including crosses involving P. communis and P. ussuriensis
(Golisz et al. 1971), P. ussuriensis crossed with P. pyrifolia and P. ×bretschneideri (Pu et al.
1963) and P. communis cultivars (Zielinski et al. 1965). Thompson et al. (1974) suggested
that the control of grit content is complex, regulated by a minimum of four loci that act in an
independent and additive fashion. In most hybrids derived from crosses of Asian and
European pears, it is difficult to recover pure crisp flesh selections with little grittiness (Wang
1990, Bell et al. 1996).
Several genetic studies have been conducted on skin and flesh color in pears. Zielinski et al.
(1965) found that the background color of pears was under the influence of a major gene,
with yellow dominant to green. They also found that the blushed vs. nonblushed character
was controlled by a recessive gene, and white flesh was dominant to colored, with green and
cream likely being regulated by other alleles at the same locus. Deep red colored sports of
'Bartlett' called 'Cardinal Red' and 'Max Red' were found to be regulated by a single dominant
gene, C by Zielinski (1963). Brown (1966) found that red was dominant to white in one
segregating population of 'Sanquinole' (red) x 'Conference' (white). Not all red skin mutations
are transmitted sexually, as the red coloration of 'Starkrimson' only affects the epidermis and
not the germlines (Dayton 1966).
Russeting of the skin is acceptable if it is smooth, uniform and light tan for the fresh market,
but is not acceptable in processed puree. In P. communis, Zielinski et al. (1965) and Crane
and Lewis (1949) found that the control of russeting was under quantitative control. Bell and
Janick (1990) estimated heritability for russeting to be 0.52. Wellington (1913) originally
suggested that russeting in P. pyrifolia was under the control of a single gene; however,
Kikuchi (1930) proposed that two loci were involved, R and I. In his model, RR__ genotypes
are completely russeted and Rrii genotypes are always partially russetted. RrI_ genotypes
have a partial russetting that is environmentally sensitive (under humid conditions they have
more russetting than under dry conditions). Wang and Wei (1987) suggested that russetting in
P. pyrifolia was recessive to non-russetting.
There is considerable variability for fruit size in Pyrus. Many genotypes of P. betulifolia and
P. calleryana are as small as 1 cm in diameter, compared to genotypes of P. communis and P.
pyrifolia which can exceed 12 cm in diameter (Bell et al. 1996). The size of pears is under
polygenic control with variable levels of heritability depending on breeding population
(Zielinski et al. 1965, Shen et al. 1979, Wang and Wei 1987, White and Alspach 1996).
White et al. (2000) found heritability for length and width ratios to be more than 0.5 in
families of European and Asian pear parentage, while Machida and Kozaki (1976) and Shin
16
et al. (1983) found much lower values in their P. pyrifolia breeding populations. Many
environmental factors influence pear size including water availability, fruit set and yield.
The most common shape of European pear cultivars is pyriform and most Asian pears are
round (Bell et al. 1996). The inheritance of fruit shape is under polygenic control (Crane and
Lewis 1949, Zielinski et al. 1965). Round and obovate shapes are most common in
segregating populations of European and Asian types, suggesting that this shape is dominant
to pyriform and turbinate shapes (Zielinski et al. 1965, Wang and Wei 1987). Shin et al.
(1983) found that the ratio of fruit length to diameter had a moderately low heritability of
0.23.
Several molecular studies have been conducted on the expression patterns of genes during
fruit ripening and storage. El-sharkawy et al. (2004) found several 1-Aminocyclopropane-1-
carboxylic acid (ACC) synthase genes in late ripening pears that were differentially expressed
during cold treatment, and they discovered that cold dependent and independent cultivars had
variant allelic assemblages. Fonseca et al. (2005) followed the transcript accumulation of
seven genes encoding cell wall modifying enzymes during fruit growth, ripening and
senescence. They found that induction of the genes for xyloglucan
endotransglucosylase/hydrolase and expansin 2 was likely associated with cell wall
maintenance, while expansin 1, polygalacturonase 1 and 2, ß-galactosidase and ß-xylosidase
most likely played a role in cell wall disassembly and loosening. Itai et al. (2000) isolated 30
cDNA clones of genes corresponding to mRNAs up-regulated during fruit ripening of
Japanese pear. The cDNAs were sequenced and were found to be associated with stress
response, protein catabolism and pathogenesis. Several of the genes were inhibited by 1-
methylcyclopropane (MCP), an inhibitor of ethylene action.
A number of genes have been isolated from pear which are associated with fruit development
and/or ripening. Lelievre et al. (1997) cloned ACC synthase and ACC oxidase from P.
communis. El-Sharkawy et al. (2003) isolated and characterized four ethylene perception
elements and found them to be differentially expressed after cold and ethylene treatment. Itai
et al. (1999a and b) characterized ACC synthase and ß-D-xylosidase from P. pyrifolia, the
latter likely playing a role in senescence. They found an association between the expression
of two specific ACC synthase genes and ethylene production in 35 Asian pear cultivars. Itai
et al. (2003) developed a rapid method for analyzing fruit storage potential by utilizing CAPS
(cleaved-amplified polymorphic sequences) markers of the two ACC synthase genes. One
marker was associated with high ethylene producers, while a second marker was associated
with moderate ethylene producers. Low ethylene producers had neither of these markers.
Sekine et al. (2006) cloned thirteen cDNAs that encode cell-wall hydrolases in the European
pear and followed their expression patterns during cold storage. Hiwasa et al. (2003) found
seven α-expansin genes to be differentially expressed during growth and ripening of pear
fruit. They identified seven genes that might be associated with the melting texture. Yamada
et al. (2006) cloned two isoforms of soluble acid invertase of Japanese pear and followed
their development during ripening; one was particularly active in young fruit. Tateishi et al.
(2001) isolated a cDNA fragment of the fruit softening enzyme, ß-galactosidase. Several
genes encoding membrane bound proteins have been cloned including arabinogalactan
proteins in P. communis (Mau et al. 1995) and H(+)-pyrophosphatase in P. pyrifolia (Suzuki
et al. 1999). Additional genes that have been characterized in pear include alcohol
17
dehydrogenase (Chervin et al. 1999), polyphenol oxidase (Haruta et al. 1999) and α-L-
arabinofuranosidase (Tateishi et al. 2005).
10.5.4 Rootstocks
Historically, little breeding effort has been devoted to improving pear rootstocks, although
Chevreau and Bell (2006) suggest that “deliberate evaluation and selection of parents and
hybridization has become more common”. Pyrus communis is the species most widely
employed as rootstocks in North America and Europe. In Asia, P. pyrifolia, P. betulifolia, P.
calleryana and P. pashia provide the primary rootstocks, while in Asia Minor and the
Mediterranean region, P. elaeagrifolia, P. syriaca, P. amygdaliformis and P. longipes are the
species most commonly utilized.
Pyrus communis seedlings produce vigorous trees that are adapted to a broad range of
climates and soil types, and they are generally resistant to pear decline and Armillaria root rot
[Armillaria mellea (Vahl:Fr.) P. Kumm.], but are often susceptible to fire blight (Bell et al.
1996). Fire blight resistant types have been found in progeny of the P. communis cross, 'Old
Home' x 'Farmingdale' (OH F rootstock series) and individuals from this family provide some
size reduction (Brooks 1984, Lombard and Westwood 1987) and precocity, but sometimes
have problems with suckering (Raese 1994). Other dwarfing P. communis rootstocks include
'Pyrodwarf' and 'BU 2-33' coming from 'Old Home' x 'Bonne Louise d'Avranches' (Jacob
1998 and 2002), the Blossier series (Brossier 1977, Michelesi 1990) and the Rètuziére series
(Michelesi 1990, Simard and Michelesi 2002). Seedlings of 'Barlett' predominate as
rootstocks in older plantings, with seedlings of 'Winter Nelis' probably being the next most
widely used seedling rootstock.
Among the other species utilized as rootstocks, Pyrus calleryana has been used in the
southern U.S.A., China and Australia because of its fire blight, pear decline and wooly aphid
resistance, drought tolerance and it’s control of vigor, although the species is very susceptible
to winter injury (Cole 1966, Batjer et al. 1967, Lombard and Westwood 1987). Pyrus
betulifolia is used in local areas of the United States, central China, northern Italy and Israel
where temperatures are not very cold; it is particularly useful where clay soils and poor
drainage restrict vigor and it is resistant to salinity, black end, pear decline and the wholly
pear aphid (Lombard and Westwood 1987, Matsumoto et al. 2006). Dwarfing root stocks
have been selected from both P. calleryana and P. betulifolia (Robbani et al. 2006). Pyrus
ussuriensis has proven useful where extreme cold hardiness is required in northeast China
and the Great Plains of North America (Morrison 1965). It is resistant to fire blight (i.e.
Reimer selections) and the woolly pear aphid, but is sensitive to excessive moisture, black
end and pear decline. Pyrus pyrifolia is utilized widely in Japan, China and Korea because of
its tolerance to a wide range of soil textures and soil moistures, even though it is prone to the
physiological disorder black end (hard end) and pear decline (Bell et al. 1996). Pyrus pashia
is used in northern India and southern China; it is resistant to black end, although it is
sensitive to lime-induced chlorosis and is susceptible to the woolly pear aphid (Bell et al.
1996). Pyrus amygdaliformis has high tolerance to salinity stress (Matsumoto et al. 2006).
Cydonia oblonga L. has long been used in Europe as a powerful dwarfing rootstock for
milder climate pear production. It reduces size by 30 to 60% compared to standard P.
communis seedling rootstocks, shortens time to fruiting and increases fruit size. However,
quince rootstocks suffer from several problems including susceptibity to fire blight and
18
Armillaria root rot, poor winter hardiness, low tolerance to wet soils, insufficient soil
anchorage and poor graft compatibity with many common pear cultivars (Millikan and
Pieniazek 1967, Lombard and Westwood 1987). Quince selections are being developed with
improved cold hardiness and greater tolerance to high pH (Loreti 1994, Bassi et al. 1996,
Webster 1998).
In addition to Cydonia, selected clones of Amelanchier and Crataegus possess graft-
compatibility with Pyrus and can be used as dwarfing rootstocks (Lombard and Westwood
1987, Lombard 1989).
10.6 Crossing and evaluation techniques
10.6.1 Breeding systems
As mentioned previously, the pear is likely an ancient allopolyploid but it behaves as a
diploid with disomic inheritance (Crane and Lewis 1940). Most Pyrus cultivars are diploid,
although there are polyploid cultivars, especially in P. communis (Bell et al. 1996). A number
of cultivars are triploid (2n = 3x = 51), a few are tetraploid (2n = 4x = 68), and hexaploid
forms have been produced, but not commercialized (2n = 6x =102).
All pear species are self-infertile with the gametophytic self-incompatibility system (Crane
and Lewis 1942, Westwood and Bjornstad 1971). Therefore, straight selfing to increase
homozygosity can not be employed, although crosses where inbreeding coefficients do not
exceed 0.25, rarely result in significant losses of vigor. A significant, albeit small, association
has been observed between inbreeding and improved flavor, grit, and texture, suggesting that
limited inbreeding can aid in selection for homozygous recessive genotypes (Bell et al. 1981).
A considerable amount of work has been done to characterize the genes associated with the
incompatibility locus (S) locus in European (Sanzol et al. 2006) and Asian pears (Sassa and
Hirano 1997, Ushijima et al. 1998, Ishimizu et al. 1999, Kim et al. 2004). PCR-RFLP and
CAPS (cleaved amplified polymorphic sequences) marker systems have been developed to
genotype the S alleles of cultivars (Kim et al. 2002, Takasaki et al. 2006, Moriya et al. 2007).
One mutated gene of S4-RNase has been described that confers self-compatibility in 'Osa-
Nijiisseiki' (Norioka et al. 1996, Wu et al. 2007). Two other genes have been cloned in pear
that are associated with pollination - the gene for uridine diphosphate (UDP)-glucose
pyrophosphorylase that plays a role in pollen tube wall synthesis (Kiyozumi et al. 1999) and a
pollen- and seed-transmitted RNA-dependent RNA polymerase (Osaki et al. 1998).
10.6.2 Pollination and seedling culture
Pollen is generally collected several weeks before the bloom period. Branches 1 to 1.5 m in
length are collected and forced in the laboratory or greenhouse with their cut ends in water.
About 2 weeks are required before the blossoms are ready for collection if the branches are
gathered while the buds are still dormant. If they are collected at the tight cluster stage, only 2
to 3 days are usually necessary.
The anthers are generally harvested just before they dehisce, by rubbing them on a wire mesh
screen over paper. The anthers are allowed to dry on the paper for 24 hours and then the
pollen is poured into glass vials that are placed in a desiccator with anhydrous CaS04 and kept
19
at about 5 °C. Pear pollen remains viable for two to three weeks at room temperature, but can
be stored for over a year with refrigeration. The pollen is generally transported to the field on
ice in larger vials containing desiccant. The vials of pollen are removed from the cooler when
required for pollination and put back immediately after use.
Emasculation is done using a variety of tools including fingernails, scalpels, tweezers, or
scissors specially modified with a notch in the blades and an adjustable screw to control the
amount of closure (Bell et al. 1996). A cut is made below the sepals and the flower is pulled
off leaving the pistil. One to three blossoms per cluster are generally emasculated at the
balloon stage of development and all other flowers are removed. The larger basal flowers
tend to set more fruit than the smaller terminal ones. Pollen is applied to the stigmas using a
variety of objects including glass rods, fingers or camel hair brushes. Bell et al. (1996)
suggests that approximately two flowers must be emasculated and pollinated to produce one
seed. Bees do not visit flowers without corollas (Visser 1951), so special precautions to
prevent pollen contamination are not necessary for routine crosses.
Pear seeds require stratification to overcome their internal dormancy requirement. This is
commonly done by holding them for 60 to 90 days in moist, finely ground peat moss placed
in polyethylene bags at temperatures a little above freezing (Hartman et al. 1990). Fungicides
are often added to the water to prevent damping off. After stratification, the seeds are planted
into flats or small pots in a sterilized mixture of equal parts sand and peat moss. Germination
generally occurs within 9 to 10 days, and the seedlings are transferred to larger pots filled
with commercial soil mix when they are 15 cm tall.
10.6.3 Evaluation techniques
Pear seedlings have a long juvenile period and remain unfruitful for 4 years or longer (Visser
et al. 1976, Bell 1991). The length of this juvenile period appears to be dependent on how
rapidly the seedlings grow (Zimmerman 1972). Therefore, growing conditions are generally
optimized as much as possible. Other factors have been evaluated to hasten flowering such as
grafting onto bearing trees or quince rootstocks, root pruning and sprays with growth
retardants, but these practices have proved generally ineffective (Zimmerman 1972,
Verhaegh et al. 1988).
There are some seedling characters, such as disease and insect resistance, that can be
conducted in the greenhouse to allow for early seedling selection (Dayton et al. 1983). These
preliminary screens can be followed with an additional screening in the nursery at close
spacing, to verify resistance ratings and eliminate escapes. Some juvenile characters are
correlated with economically important adult characteristics. Bell et al. (1996) list some of
these in P. communis such us seedling vigor with precocity (Zimmerman 1972), juvenile
period with precocity of propagated trees (Visser and De Vries 1970), juvenile period with
red leaf coloration (Zielinski 1963), early flowering before leaf break with high yield (Moruju
and Slusanschi 1959), seed size with germinability and seedling vigor (Schander 1955), and
fruit color with foliage color.
When the plants do fruit, selection is undertaken on such characters as precocity and
productivity, fruit quality, harvest dates and ripening uniformity. The storage and processing
qualities of the most promising types may also be evaluated if there is enough fruit. The
selected genotypes are then evaluated as grafted trees at normal orchard spacing in one
20
location, and then the most promising are further evaluated as grafted trees at multiple
locations.
10.7 Biotechnological approaches to genetic improvement
10.7.1 Genetic mapping and QTL analysis
A wide array of molecular markers have been developed for pear including isozymes,
restriction fragment length polymorphisms (RFLPs), randomly amplified polymorphic DNA
(RAPDs), amplified fragment length polymorphisms (AFLPs) and inter simple sequence
repeats (ISSRs) (Chevreau and Bell 2006). This work has primarily focused on fingerprinting
cultivars and measuring diversity patterns within and among Pyrus species (Monte-Corvo et
al. 2000, Yamamoto et al. 2002a and b, Inoue et al. 2007, Katayama and Uematsu 2006). In
general, cultivars have been readily distinguished through these approaches and species have
clustered according to traditional taxonomic classifications. Many SSRs isolated from apple
have been shown to be transferable to pear (Yamamoto et al. 2001 and 2002c, Pierantoni et
al. 2004), and vice versa (Fernández-Fernández et al. 2006).
Physical maps of chloroplast DNA (cpDNA) have also been generated using restriction
analysis. In a comparison of Asian and Occidental pear species, Iketani et al. (1998) found
four cpDNA haplotypes in Asian pears, but only one in occidental pears. They argued that the
two groups have evolved separately, and that considerable hybridization and introgression
has occurred between species. Katayama and Uematsu (2003) developed a cpDNA map of
Pyrus ussuriensis var. hondoensis and then carried out an RFLP analysis on cpDNAs from
representatives of Pyrus pyrifolia, P. ussuriensis, P. calleryana, P. elaeagrifolia and P.
communis. Two mutations, a recognition-site mutation and a length mutation (deletion), were
found in the cpDNA of P. pyrifolia cultivars.
Several genetic linkage maps have been developed of pear. Isozyme loci were used to
identify three linkage groups in pear that shared a high level of synteny with apple (Chevreau
et al. 1997). Iketani et al. (2001) employed RAPD markers to construct linkage groups of 18
and 22 for the two Asian pears, 'Kinchaku' and 'Kosui'. The linkage map for 'Kinchaku'
contained 120 loci in 18 linkage groups across 768 cM, while that for 'Kosui' had 78 loci in
22 linkage groups spanning over 508 cM.
Yamamoto et al. (2002c) constructed a genetic linkage map of an interspecific cross between
European (P. communis cv. 'Bartlett') and Asian (P. pyrifolia cv. 'Housui') pears using
isozymes, AFLPs, SSRs and morphological traits from pear, apple, peach and cherry. In the
map of the female parent, 'Bartlett', 226 loci were identified on 18 linkage groups over a total
length of 926 cM. In the male parent, 'Housui', 154 loci were represented on 17 linkage
groups encompassing a genetic distance of 926 cM. The position of 14 SSRs from apple
could be placed on the pear map, along with a few SSRs from Prunus.
Pierantoni et al. (2004) used 100 apple SSRs to develop linkage maps of two European pear
families. A total of 41 markers were positioned on the cross of 'Passe Crassane' x 'Harrow
Sweet', and 31 were placed on a map of 'Abbè Fetél' x 'Max Red Bartlett'. Considerable
colinearity was observed in the linkage relationships of the apple and pear genomes. Dondini
et al. (2004) expanded the map of 'Passe Crassane' x 'Harrow Sweet' with a wide array of
markers including SSRs, MFLPs (microsatellite-anchored fragment length polymorphisms),
21
AFLPs and RGAs (resistance gene analogs). They placed 155 loci on the 'Passe Crassane'
map consisting of 18 linkage groups with a coverage of 912 cM. On the Harrow Sweet map
they identified 156 loci on 19 linkage groups with a coverage of 930 cM.
Figure 10.3. QTL for fire blight resistance identified on the linkage groups of the fire blight
tolerant cultivar 'Harrow Sweet'. The probability of association of markers is indicated by the
LOD score. The black line indicates a LOD threshold value of 1.3, which is the 95%
probability value (Dondini et al. 2005).
DNA markers have been linked to a number of genes of horticultural importance in pear.
Banno et al. (1999) identified a RAPD marker that was closely linked to the gene A,
conferring susceptibility to black spot disease in Asian pear. Inoue et al. (2006) found a
RAPD marker linked to fruit skin color in Japanese pear. Iketani et al. (2001) found four
RAPD markers that were loosely associated with the pear scab resistance gene, Vnk; they
were able to place the resistance allele for pear scab and the susceptible gene for black spot
on their linkage map. Terakami et al. (2006) also found an SSR marker closely linked to the
22
scab resistance gene, Vnk, and converted the sequence into a sequence tagged site (STS)
marker, along with the four RAPD markers previously identified by Iketani et al. (2001).
Dondini et al. (2004) found four putative QTL for fire blight resistance (Figure 10.3).
10.7.2 Somatic cell genetics and genetic manipulation
Haplo-diploidization through gametic embryogenesis has been employed in pears to obtain
homozygous lines from heterozygous parents (Germaná 2006). Bouvier et al. (1993) found
haploid plants (2n = x = 17) among seedlings of 12 crosses of European pear and were able to
induce in situ parthenogenesis using irradiated pollen. The immature embryos were cultured
in vitro and 1 haploid, two misoploid and several diploids with the maternal phenotype were
recovered. In subsequent work, Bouvier et al. (2002) treated haploids with oryzalin in vitro to
generate doubled haploids. These were confirmed to be diploid and homozygous using
isozyme and microsatellite markers. Kadotat and Niimi (2002 and 2004) have produced
triploid plants of Japanese pear by anther culture, and tetraploids using in vitro colchicine
treatment.
Regeneration techniques have been developed for a wide array of elite pear cultivars and
genotypes, including those of P. communis (Chevreau et al. 1997, Matsuda et al. 2005,
Yancheva et al. 2006), P. pyrifolia (Lane et al. 1998); P. syriaca (Shibli et al. 2000), P.
pyraster (Caboni et al. 1999) and quince (Dolcet-Sanjuan et al. 1991, Baker and Bhatia
1993). Leaves from in vitro grown plants were used as explants in most work, with a brief
amount of callus growth at the wounding site before bud regeneration in about 3-6 weeks.
Explants were generally exposed to dark and light periods of 2 – 4 weeks. The most common
hormones employed were thidiazuron (TDZ) and naphthaleneacetic acid (NAA). Maximum
rates of regeneration varied from < 20% in P. pyrifolia (Lane et al. 1998) to > 80 % in P.
communis (Chevreau and Leblay 1993) and C. oblonga (Baker and Bhatia 1993).
Palombi et al. (2007) used in vitro regeneration of wild pear (P. pyraster) to generate
somaclonal variants for higher adaptability to calcareous soils. Selective treatments involved
Murashige and Skoog (MS) medium with Fe-EDTA replaced by equimolar amount of FeSO4
with KHCO3 or NaHCO3. Eleven putatively tolerant lines were obtained from vegetatative
shoot apices.
The first transformed pears were the European cultivars 'Passe Crassane', 'Conference' and
'Doyenne du Comice' (Mourgues et al. 1996). Since then, several other cultivars of P.
communis have been transformed including 'Vyzhnitsa' (Merkulov et al. 1998), 'Beurre Bosc'
(Bell et al. 1999), 'Barlett' (Bommineni et al. 2001), wild P. pyraster (Caboni et al. 1998), P.
betulifolia seedlings (Kaneyoshi et al. 2001) and the rootstocks GP217 (Lebedev and Dolgov
2000) and BP100030 (Zhu and Welander 2000). A number of different disarmed strains of
Agrobacterium tumifaciens were utilized in these studies, with kanamycin or hygromycin
being used as the selectable markers. In most cases, leaves were used as explants, with the
only exception being cotyledons of P. betulifolia.
Transformation strategies have been employed in a number of instances to improve disease
resistance. Researchers at INRA in Angers, France, have incorporated transgenes encoding
lytic peptides from insects (attacin and cecropin) (Reynoird et al. 1999), lysozymes from T4
bacteriophage (Mourgues et al. 1998), the lactoferrin gene from bovin (Malnoy et al. 2003a)
and a viral EPS-depolymerase gene (Malnoy et al. 2002) to develop resistance to fire blight.
23
The insertion of the lytic peptide gene (D5C1) was shown to be partially effective against
pear psylla (Puterka et al. 2002). Lededev et al. (2002) introduced plant defensin genes into
pear to enhance pathogen resistance. Two genes have been cloned that may play a role in
defense against pathogens, polygalacturonase inhibitor protein (PGIP) (Stotz et al. 1993) and
thaumatin/PR5-like protein (Sassa and Hirano 1998). Malnoy et al. (2003b) searched for
pathogen-inducible promoters from tobacco that would work in pear and found two, str246C
and sgd24, which were responsive to inoculation by Erwinia amylovora.
In other transformation work, the phophinotricin acetyl transferase (PAT) gene was
incorporated into pear rootstocks to produce herbicide resistance (Lebedev et al. 2002b), and
the super sweet gene thaumatin II was used to modify fruit taste (Lebedev et al. 2002c). The
rolC gene from Agrobacterium rhizogenes was introduced into pear to cause dwarfing (Bell
et al. 1999) and enhance rooting (Zhu and Welander 2000). The gene encoding S-
adenosylmethionine hydrolase (sam-k) was incorporated into 'Bartlett' to improve postharvest
quality and self-life by modifying ethylene synthesis (Bommineni et al. 2000).
10.7.3 Mutation breeding
Irradiation (X-rays) has been used to increase the frequency of mutations in fruit trees
(Ahloowalia et al. 2004). Several kinds of mutations have been identified after irradiation in
P. communis including bloom time, blossom color, ripening time, fruit color (Decourtye
1970, Roby 1972a and b, Predieri and Zimmerman 2001) and growth habit (Visser et al.
1971, Lacey 1975, Predieri and Zimmerman 1997). In P. pyrifolia, mutations have been
induced that effected disease resistance (Matuda et al. 1997) and self-compatibility (Hirata
1989). At least five European and four Japanese pears have been developed through mutation
breeding. One of them, 'Gold Nijisseiki', has had a substantial impact on the Asian pear
industry, and two new self-compatible, black spot resistant varieties show high promise
(Ahloowalia et al. 2004).
10.8 References
Ahloowalia BS, Maluszynski M, Nichterlein K (2004) Global impact of mutation-derived varieties. Euphytica
135:187-204
Anjou K (1954) Winter injury of apples and pears at Balsgard, 1953. (in Swedish). Sverig Pomol Foren Arsskr
54:139-147
Baker BS, Bhatia SK (1993) Factors affecting shoot regeneration from leaf explants of quince (Cydonia
oblonga). Plant Cell Tiss Org Culture 35:273-277
Baldini E (1949) Spring frost damage to fruit trees in the spring of 1949 (in Italian). Riv Ortoflorofruttic Ital
33:78-86
Banno K, Ishikawa H, Hamauzu Y, Tabira H (1999) Identification of a RAPD marker linked to the susceptible
gene of black spot disease in Japanese pear. J Jap Soc Hort Sci 68:476-481
Bao L, Chen K, Zhang D, Cao Y, Yamamoto T, Teng Y (2007) Genetic diversity and similarity of pear (Pyrus
L.) cultivars native to East Asia revealed by SSR (simple sequence repeat) markers. Genet Resour Crop
Evol 54:959-971
Barbosa W, Pommer CV, Tombolato AFC, Meletti LMM, Veiga RF de A, Moura MF, Pio R (2007) Asian pear
tree breeding for subtropical areas of Brazil. Fruits-Paris 62: 21-26
Bassi D, Marangoni B, Tagliavini M (1996) ‘Fox’, nuova serie di portinnest clonai per il pero. Frutticoltura
3:55-59
Batjer LP, Schomer HA, Newcomer EJ, Coyier DL (1967) Commercial pear growing. USDA Handb. 330
Bell RL (1991) Pears (Pyrus). Acta Hort 290:657-700
Bell RL (1992) Additionall East European Pyrus germplasm with resistance to pear psylla nymphal feeding.
HortScience 27:412-413
Bell RL, Hough LF (1986) Interspecific and intergeneric hybridization of Pyrus. HortScience 21:62-64
24
Bell RL, Janick J (1990) Quantitative genetic analysis of fruit quality in pear. J Am Soc Hort Sci 115:829-834
Bell RL, Quamme HA, Layne REC, Skirvin RM (1996) Pears In: Janick J, Moore JN (eds.) Fruit Breeding Vol
1. Tree and Tropical Fruits. John Wiley and Sons, NY, pp. 441-514
Bell RL, Scorza R, Srinivasan C, Webb K (1999) Transformation of ‘Beurre Bosc’ pear with the rolC gene. J
Amer Soc Hort Sci 124:570-574
Bell RL, Stuart LC (1990) Resistance in Eastern European Pyrus germplasm to pear psylla nymphal feeding.
HortScience 25:789-791
Bell RL, van der Zwet T (2005) Host resistance in Pyrus to Fabraea leaf spot. HortScience 40:21-23
Bell RL, van der Zwet T, Blake RC (2002) ‘Blake’s Pride’ pear. HortScience 37:711-713
Bell RL, Zimmerman RH (1990) Combining ability analysis of juvenile period in pear. HortScience 25:1425-
1427
Bell RL, Janick J, Zimmerman RH, van der Zwet T (1977) Estimation of heritability and combining ability for
fire blight resistance in pear. J Amer Soc Hort Sci 102:133-138
Bell RL, Janick J, Zimmerman RH, van der Zwet T, Blake RC (1981) Response of pear to inbreeding. J Amer
Soc Hort Sci 106:584-589
Bellini E, Sansavini S, Lugli S, Nin S, Rivlta L (2000) Obiettivi innovatori del miglioramento genetico del pero.
Rivista Frutt Ortoflor 62:56-69
Bommineni VR, Mathews H, Clenennen SK, Wagoner W, Dewey V, Kellogg J, Peters S, Matsumura W, Pieper
M, Kramer M, Wagner DR (2000) Genetic engineering of fruits and vegetables with the ethylene
control gene encoding S-adenosylmethionine hydrase (SAMase). Proceedings Inter Sym Plant Genetic
Engineering, Havana, Cuba. pp 206-214
Bommineni VR, Mathews H, Samuel SB, Kramer M, Wagner DR (2001) A new method for rapid in vitro
propogation of apple and pear. HortScience 36:1102-1106
Bouvier L, Guérif P, Djulbic M, Durel CE, Chevreau E, Lespinasse Y (2002) Chromosome doubling of pear
haploid plants and homozygosity assessment using isozyme and microsatellite markers. Euphytica
123:255-262
Bouvier L, Zhang Y-X, Lespinasse Y (1993) Two methods of haploidization in pear, Pyrus communis L.:
Greenhouse seedling selection and in situ parthenogenesis induced by irradiated pollen. Theor Appl
Genet 87:229-232
Braniste N (2002) The pear industry in Eastern Europe. Acta Hort 596:83-85
Brooks L (1984) History of the Old Home x Farmingdale pear rootstocks. Fruit Var J 38:126-128
Brown AG (1960) Scab resistance in progenies of varieties of the cultivated pear. Euphytica 9:247-253
Brown AG (1966) Genetical studies in pears, 5: red mutants. Euphytica 15:425-429
Budagovskij VI (1966) The use of artificial methods of freezing roots for determining the frost resistance of
rootstocks. 37-43. In: Trushechkin VG, Tarakanov GI, and Nicolaenko NP (eds.), Reports of the Soviet
Scientist to the 17th Int. Congress on Horticulture. Ministry of Agriculture of the USSR, Moscow
Caboni E, Tonelli MG, Lauri P, D’Angeli S, Damiano C (1999) In vitro shoot regeneration from leaves of wild
pear. Plant Cell Tiss Organ Cult 59:1-7
Caboni E, Tonelli MG, Lauri P, Damiano C (1998) Regeneration and transformation of pear rootstocks. In:
Abstracts XXV Inter Hort Cong, August 2-7, Brussels
Challice J, Westwood MN (1973) Numerical taxonomic studies of the genus Pyrus using both chemical and
botanical characteristics. Bot J Linn Soc 67:121-148
Chervin C, Spiers J, Truett J (1999) Alcohol dehydrogense expression and the production of alcohols during
pear fruit ripening. J Amer Soc Hort Sci 124:71-75
Chevreau E, Bell R (2004) Pyrus spp. pear and Cydonia spp. Quince. P. 543-565. Litz RE (ed) Biotechnology of
fruit and nut crops. CABI Publishing, Wallingford, UK
Chevreau E, Leblay C (1993) The effect of mother plant pretreatment and explant choice on regeneration from
in vitro pear leaves. Acta Hort 336:263-268
Chevreau E, Leuliette S, Gallet M (1997) Inheritance and linkage of isozyme loci in pear (Pyrus communis L.).
Theor Appl Genet 94:498-506
Cole CE (1966) The fruit industry of Australia and New Zealand. Proc 17th Int Hor Congr 3:321-368
Crane MB, Lewis D (1940) Genetical studies in pears, 2: classification of culture varieties. J Pomol 18:52-60
Crane MB, Lewis D (1942) Genetical studies in pears, 3: incompatibility and sterility. J Genet. 43:31-43
Crane MB, Lewis D (1949) Genetical studies in pears, 5: vegetative and fruit characters. Heredity 3:85-97
Dayton DF (1966) The pattern and inheritance of anthocyanin distribution in red pears. Proc Amer Soc Hort Soc
89:110-116
Dayton DF, Bell RL, Williams EB (1983) Disease resistance breeding. p 119-159. In: Moore JN, Janick J (eds.)
Methods in fruit breeding. Purdue University Press, West Layfayette, IN
25
Decourtye L (1967) Etude de quelques caracteres a controle genetique simple chez le pommier (Malus sp.) et le
poirier (Pyrus communis L.). Ann Amel Plantes 17:243-265
Dolcet-Sanjuan R, Mok DWS, Mok MC (1991) Plantlet regeneration from cultured leaves of Cydonia oblonga
L. (quince). Plant Cell Rep 10:240-242
Dondini L, Pierantoni L, Gaiotti F, Chiodini R, Tartarini S, Bazzi C, Sansavina S (2004) Identifying QTLs for
fire blight resistance via European pear (Pyrus communis L.) genetic linkage map. Mol Breeding
14:407-418
El- Sharkawy I, Jones B, Gentzbittel L, Lelièvre J-M, Pech JC, Latché A (2004) Differential regulation of ACC
synthase genes in cold-dependent and –independent ripening in pear fruit. Plant Cell Environ 27:1197-
1210
El- Sharkawy I, Jones B, Li ZG, Lelièvre J-M, Pech JC, Latché A (2003) Isolation and characterization of four
ethylene perception elements and their expression during ripening in pears (Pyrus communis L.)
with/without cold requirement. J Exper Bot 54:1615-1625
Fernández-Fernández F, Harvey NG, James CM (2006) Isolation and characterization of polymorphic
microsatellite markers from European pear (Pyrus communis L.). Mol Eco Notes 6:1039-1041
Fisher DF (1922) An outbreak of powdery mildew (Podosphaera leucotricha) on pears. Phytopathology 12:103
Fisher M, Mildenberger (2000) New Naumburg/Pillnitz pear breeding results. Acta Hort 538:735-739
Fonseca S, Monteiro L, Barreiro MG, Pais M (2005) Expression of genes encoding cell wall modifying enzymes
is induced by cold storage and reflects changes in pear fruit texture. J Exp Bot 56:2029-2036
Germaná MA (2006) Doubled haploid production in fruit crops. Plant Cell Tis Org Cult 86:131-146
Ghariani K, Stebbins RL (1994) Chilling requirements of apple and pear cultivars. Fruit Var J 48:215-222
Golisz A, Basak A, Zagaja SW (1971) Pear cultivar breeding. In: S. A. Pieniazd (ed.) Studies on some local
Polish fruit species, varieties and clones and on thost recently introduced to Poland with respect to their
breeding value and other characters Nov. 1, 1966, to Oct. 31, 1971. Res Inst Pomology, Skierniewice,
Poland
Harris MK, Lamb RC (1973) Resistance to the pear psylla in pears with Pyrus ussuriensis lineage. J Amer Soc
Hort Sci 98:378-381
Hartman H (1957) Catalog and evaluation of the pear collection at the Oregon Agriculture Experiment Station.
Oregon Agric Exp Sta Tech Bull 41
Hartmann HT, Kester DE, Davies FT (1990) Plant propagation: principles and practices, 5th ed. Prentice-Hall,
Englewood Cliffs, NJ
Haruta M, Murata M, Kadokura H, Homma S (1999) Immunological and molecular comparison of polyphenol
oxidase in Rosaceae fruit trees. Phytochem 50:1021-1025
Hedrick UP, Howe GH, Taylor OM, Francis EH, Tukey HB (1921) The pears of New York. New York Dept.
Agric. 29th Annu Rep Vol 2, part 2
Hevesi M, Göndör M, Kása K, Honty K, Tóth MG (2004) Traditional and commercial apple and pear cultivars
as sources of resistance to fireblight. EPPO/OEPP Bull 34:377-380
Hibino H, Schneider H (1970) Mycoplasmalike bodies in sieve tubes of pear trees affected with pear decline.
Phytopathology 60:499-501
Hiroe I, Nishimura S, Sato M (1958) Pathochemical studies on Alternaria kikuchiana: on toxins secreted by the
fungus (in Japanese). Trans Tottori Soc Agr Sci 11:291-299
Hirata N (1989) Self-compatible mutant in Japanese pear. Gamma Field Symp 28:71-80
Hiwasa K, Rose JKC, Nakano R, Inaba A, Kubo Y (2003) Differential expression of seven α-expansin genes
during growth and development. Physiol Plant 117:564-572
Hooker W (1818) Account of new pear, called Williams' Bon Chretien. Trans Hort Soc London 2:250-251
Horvat RJ, Senter SD, Chapman GW, Payne JA (1992) Volatiles of ripe Asian pears (Pyrus serotina Rehder). J
Essent Oil Res 4:645-646
Hsu H-T, Lin S-C (1987) Oriental pear breeding for high quality and adaptation to subtropical lowlands of
Taiwan. p. 95-100. In: The breeding of horticultural crops. Food and Fertilizer Technology Center for
the Asian and Pacific Region, Taipei
Human JP (2005) Progress and challenges of the South African pear breeding. Acta Hort 671:185-190
Hunter DM, Kappel F, Quamme HA, Bonn WG (2002a) ‘AC Harrow Gold’ pear. HortScience 37:224-226
Hunter DM, Kappel F, Quamme HA, Bonn WG (2002b) ‘AC Harrow Crisp’ pear. HortScience 37:227-229
Jennings WG, Creveling RK, Heintz DE (1964) Volatile esters of Bartlett pear. IV. Esters of Trans:2-cis:4-
decadienoic acid. J Food Chem 29:730-734
Iketani H, Abe K, Yamamoto T, Kotobuki K, Sato Y, Saito T, Terai O, Matsuta N, Hayashi T (2001) Mapping
of disease-related genes in Japanese pear using a molecular linkage map with RAPD markers. Breeding
Sci 51:179-184
26
Iketani H, Manabe T, Matsuta N, Akihama T, Hayashi T (1998) Incongruence between RFLPs of chloroplast
DNA and morphological classicfication in east Asian pear (Pyrus spp.). Genet Res Crop Evol 45:533-
539
Inoue E, Kasumi M, Sakuma F, Anzai H, Amano K, Hara H (2006) Identification of RAPD marker linked to
fruit skin color in Japanese pear (Pyrus pyrifolia Makai). Sci Hort 107:254-258
Inoue E, Matsuki Y, Anzai H, Evans K (2007) Isolation and characterization of microsatellite markers in
Japanese pear (Pyrus pyrifolia Nakai). Mol Eco Notes 7:445-447
Ishimizu T, Inoue K, Shimonaka M, Saito T, Terai O, Norioka S (1999) PCR-based method for identifying the
S-genotype of Japanese pear cultivars. Theo Appl Genet 98:961-967
Itai A, Kawata T, Tanabe K, Tamura F, Uchiyama M, Tomomitsu M, Shiraiwa N (1999a) Identification of 1-
aminocyclopropane-1-carboxylic acid synthase genes controlling the ethylene level of ripening fruit in
Japanese pear (Pyrus pyrifolia Nakai). Mol Gen Genet 261:42-49
Itai A, Kotaki T, Tanabe K, Tamura F, Kawaguchi D, Fukuda M (2003) Rapid identification of 1-
aminocyclopropane-1-carboxylate (ACC) synthase geneotypes in cultivars of Japanese pear (Pyrus
pyrifolia Nakai) using CAPS markers. Theor Appl Genet 106:1266-1272
Itai A, Tanabe K, Tamura F, Tanaka T (2000) Isolation of cDNA clones cooresponding to genes expressed
during fruit ripening in Japanese pear (Pyrus pyrifolia Nakai): Involvement of the ethylene signal
transduction pathway in their expression. J Exp Bot 51:1163-1166
Itai A, Yoshida K, Tanabe K, Tamura F (1999b) A beta-D-xylosidase-like gene is expressed during fruit
ripening in Japanese pear (Pyrus pyrifolia Nakai). J Exp Bot 50:877-878
Jacob H (1998) Pyrodwarf, a new clonal rootstock for high density pear orchards. Acta Hort 475:169-177
Jacob H (2002) New pear rootstocks from Geisenheim, Germany. Acta Hort 596:337-344
Jennings WG, Tressel R (1974) Production of volatile compounds in the ripening of Bartlett pear. Chem
Mikrobiol Technol Lebensm 3:52-55
Jingxian J, Shenpu F, Lanchen C (1988) Evaluation and utilization of compact pear germplasm resource. p. 44.
In: Int. Symp. Hort. Germplasm, Cultivated and Wild. (Abstracts), Beijing, China, Sept. 1988. Int Acad
Pub, Beijing, China
Kadota M, Niimi Y (2002) In vitro induction of tetraploid plants from a diploid Japanese pear cultivar (Pyrus
pyrifolia N. vc. Hosui). Plant Cell Rep 21:282-286
Kadota M, Niimi Y (2004) Production of triploid plants of Japanese pear (Pyrus pyrifolia Nakai) by anther
culture. Euphytica 138:141-147
Kajiura I (1992) Nashi (Japanese pear) production in Japan. Chron Hort 32:57-58
Kajiura I, Sato Y (1990) Recent progress in Japanese pear (Pyrus pyrifolia Nakai) breeding, and descriptions of
cultivars based on literature review. Bull Fruit Tree Res Sta, Extra N. 1
Kajiura M (1966) The fruit industry of Japan, South Korea and Taiwan. Proc 17th Int Hort Congr 4:403-425
Kanato K, Kajiura I, McKenzie DW (1982) The ideal Japanese pear. p. 138-155. In: van der Zwet T, Childers
NF (eds.) The pear. Horticultural Publ, Gainesville, FL
Kaneyoshi J, Wabiko H, Kobayashi S, Tsuchiya T (2001) Agrobacterium tumefaciens AKE10-mediated
transformation of an asian pear, Pyrus betulaefolia Bunge: host specificity of bacterial stains. Plant
Cell Rep 20:622-628
Katayama H, Adachi S, Yamamoto T, Uematsu C (2007) A wide range of genetic diversity in pear (Pyrus
ussuriensis var. aromatica) genetic resources from Iwate, Japan revealed by SSR and chloroplast DNA
markers. Genet Resour Crop Evol (Online First)
Katayama H, Uematsu C (2003) Comparative analysis of chloroplast DNA in Pyrus species: Physical map and
gene localization. Theor Appl Genet 106:303-310
Katayama H, Uematsu C (2006) Pear (Pyrus species) genetic resources in Iwate, Japan. Genet Res Crop Evol
53:483-498
Khan KA (1955) Studies on stocks immune to woolly aphids of apple (Eriosoma lanigerum Hausm.). Punjab
Fruit J 19:28-35
Kikuchi A (1946) Speciation and taxonomy of Chinese pears. Collected Records of Hort Res 3:1-8. Kyoto Univ
Kikuchi A (1930) On skin color of the Japanese pear, and its inheritance (in Japanese). Contr Inst Plant Ind 8:1-
50
Kim H-T, Hirata Y, Nou I-S (2002) Determination of S-genotypes of pear (Pyrus pyrifolia) cultivars by S-
RNase sequencing and PCR-RFLP analysis. Mol Cells 13:444-451
Kim H-T, Hirata Y, Shin Y-U, Hwang H-S, Hwang J-H, Shin I-S, Kim D-I, Kang S-J, Kim H-J, Shin D-Y, Nou
I-S (2004) A molecular technique for selection of self-compatible varieties of Japanese pear (Pyrus
pyrifolia Nakai). Euphytica 138:73-80
27
Kim W-C, Ko K-C (1991) Studies on the inheritance of major characters in oriental pear cultivars (Pyrus
pyrifolia Nakai and P. pyrifolia x P. ussuriensis). 1. The inheritance of physiological characteristics
(colors of leaf, fruit skin, blooming time, maturity and diseases). Res Rep Rural Dev Admin 33:76-84
Kim W-C, Ko K-C (1992) Studies on the inheritance of major characters in oriental pear cultivars (Pyrus
pyrifolia Nakai and P. pyrifolia x P. ussuriensis). 2. The inheritance of characteristics concerned with
fruit shape, fruit volume and fruit weight. Res Rep Rural Dev Admin 34:51-59
Kiyozumi D, Ishimizu T, Nakanishi T, Sakiyama F, Norioka S (1999) Molecular cloning and nucleotide
sequencing of a cDNA encoding UDP-glucose pyrophosphorylase of Japanese pear (Pyrus pyrifolia).
Plant Physiol 119:364
Knight RL (1963) Abstract bibliography of fruit breeding and genetics to 1960 Malus and Pyrus. E. Malling,
Commonwealth Agric Bur Tech Commun 29
Kovalev NV (1963) Leaf blight of pears (in Russian). Zasc Rast Vred Bolez 8:58
Kozaki I (1973) Black spot disease resistance in Japanese pear. I: Inheritance of disease resistance (in Japanese).
Bull Hort Res Sta Jpn A 12:17-27
Kuznetzov PV (1941) The role of Pyrus salicifolia Pall. in the development of fruit growing in arid regions (in
Russian). Sovetsk Bot 1-2:103-107
Lacey CND (1975) Induction and selection of mutant form of fruit plants. Long Ashton Ann Rep 22-24
Lane WD, Iketani H, Hayashi T (1998) Shoot regeneration from cultured leaves of Japanese pear (Pyrus
pyrifolia) Plant Cell Tiss Organ Cult 54:9-14
Lantz HL (1929) Pear breeding: an inheritance study of Pyrus communis x P. ussuriensis hybrid fruits. Proc Am
Soc Hort Sci 26:13-19
Layne REC (1968) Breeding blight-resistant pears for southwestern Ontario. Can Agric 13:28-29
Lebedev VG, Dolgov SV (2000) The effect of selective agents and a plant intron on transformation efficiency
and expression of heterologous genes in pear Pyrus communis L. Rus J Genet 36:650-655
Lebedev VG, Lavrova N, Lunin VG, Dolgov SV (2002a) Plant-defensin genes introduction for improvement of
pear phytopathogen resistance. Acta Hort 596:167-172
Lebedev VG, Skriabin KG, Dolgov SV (2002b) Transgenic pear clonal rootstocks resistant to the herbicide
‘Basta’. Acta Hort 596:193-197
Lebedev VG, Taran SA, Dolgov SV (2002c) Pear transformation by gene of supersweet protein thaumatin II for
fruit taste modification. Acta Hort 596:199-202
Leliévre JM, Tichit L, Dao P, Fillion L, Nam YM, Pech JC, Latche A (1997) Effects of chilling on the
expression of ethylene biosynthesis genes in Passe Crassane pear (Pyrus communis L.) fruits. Plant
Mol Biol 33:847-855
Li ZL, Shen DX, Zheng SQ (1981) On the “juvenile span”, fruiting and inheritance of pear seedlings. 1-
Zhejiang Agric Univ 7(3): December
Lombard PB (1989) Dwarfing rootstocks for European pear. Compact Fruit Tree 22:74
Lombard PB, Westwood MN (1987) Pear rootstocks. p. 145-183. In: Rom RC, Carlson RF (eds.) Rootstocks for
fruit crops. Wiley, NY
Loreti F (1994) Attuali conoscenze sui principali portinnesti degli alberi da futto. Rivista Frutticoltura 9:21-26
Luby JJ, Bedford DS, Hoover EE, Munson ST, Gray WH, Wildung DK, Stushnoff C (1987) 'Surnmercrisp'
pear. HortScience 22:964
Luckwill LC, Pollard A (1963) Perry pears. Univ. Bristol Press, Bristol, UK
Ludin Y (1942) Hardiness of fruit trees in the winter of 1941-42 (in Swedish). Fruktodlaren 6:168-171
Machida Y, Kozaki I (1975) Quantitative studies on the fruit quality of Japanese pear (Pyrus serotina Rehder)
breeding, I: statistical analysis of cultivar populations (in Japanese). J Jpn Soc Hort Sci 44:235-240
Machida Y, Kozaki I (1976) Quantitative studies on the fruit quality of Japanese pear (Pyrus seratina Rehder)
breeding, II: statistical analysis of a hybrid seedling population (in Japanese). J Jpn Soc Hort Sci
44:325-329
Machida Y, Maeda M (1966) Studies on the texture of pear fruit, II: factors responsible for the texture in
Japanese pear varieties, I: fruit firmness and density of stone cell cluster (in Japanese). Bull Hort Res
Sta Jpn A5:121-129
Magness JR (1937) Progress in pear improvement. p. 615-630. In: USDA Yearb. Agr USDA, Washington, DC
Malnoy M, Brisset MN, Chevreau E (2002) Expression of a depolymerase gene in transgenic pears increased
only slightly their fire blight resistance. Acta Hort 590:401-405
Malnoy M, Venisse JS, Brisset MN, Chevreau E (2003a) Expression of bovine lactoferrin cDNA confers
resistance to Erwinia amylovora in transgenic pear. Mol Breeding 12:231-244
Malnoy M, Venisse JS, Reynoird JP, Chevreau E (2003b) Activation of three pathogen-inducible promoters of
tobacco in transgenic pears (Pyrus communis L.) after biotic and abiotic elicitation. Planta 216:802-814
Matjunin NF (1960) The question of frost resistance in fruit growing, Vol. 11 p. 31-33. Sadovodstvo, Moscow
28
Matsuda N, Gao M, Isuzugawa K, Takashina T, Nishimura K (2005) Development of an Agrobacterium-
mediated transformation method for pear (Pyrus communis L.) with leaf-section and axillary shoot-
meristem explants. Plant Cell Rep 24:45-51
Matsumoto K, Tamura F, Chun J-P, Tanabe K (2006) Native Mediterranean Pyrus rootstock, P. amygdaliformis
and P. elaeagrifolia present higher tolerance to salinity stress compared with Asian natives. J Japan
Soc Hort Sci 75:450-457
Mau SL, Chen CG, Pu ZY, Mority RL, Simpson RJ, Bacic A, Clark AE (1995) Molecular cloning of cDNAse,
coding the protein backbones of arabinogalactan-proteins from the filtrate of suspension-cultured cells
of Pyrus communis and Nicotiana alata. Plant J 8:269-281
Merkulov SM, Bartish IV, Dolgov SV, Pasternak TP, McHugen A (1998) Genetic transformation of pear Pyrus
communis L. mediated by Agrobacterium tumefaciens. Genetica Moskva 34:373-378
Michelesi JC (1990) Les porte-griffes du poirer. L’Arboriculture Fruitiere 427:19-27
Millikan DF, Pieniazek SA (1967) Superior quince rootstocks for pear from east Europe. Fruit Var Hort Dig
21:2
Monte-Corvo L, Cabrita L, Oliveira C, Leitão J (2000) Assessment of genetic relationships among Pyrus
species and cultivars using AFLP and RAPD markers. Genet Res Crop Evol 47:257-265
Moriya Y, Yamamoto K, Okada K, Iwanami H, Bessho H, Nakanishi T, Takasaki T (2007) Development of a
CAPS marker system for genotyping European pear cultivars harboring 17 S alleles. Plant Cell Rep
26:345-354
Morrison JW, Cumming WA, Temmerman HJ (1965) Tree fruits for the prairies. Can Dept Agric Publ 1222
Moruju G, Slusanschi H (1959) The study of the correlation between the processes of growth and fruiting at the
commencement of shooting in some apple and pear varieties (in Romanian). p. 317-330. In: Lucar Sti
Inst Cerut Hort-Vit Baneasa-Bucuresti, 1957
Mourgues F, Chevreau E, Lambert C, de Bondt A (1996) Efficient Agrobacterium-mediated transformation and
recovery of transgenic plants from pear (Pyrus communis L.). Plant Cell Reports 16:245-249
Mourgues F, Brisset MN, Chevreau E (1998) Activity of different antibacterial peptides on Erwinia amylovora
growth, and evaluation of the phytotoxicity and stability of cecropins. Plant Sci 139:83-91
Norioka N, Norioka S, Ohnishi Y, Ishimizu T, Oneyama C, Nakanishi T, Sakiyama F (1996) Molecular cloning
and nucleotide sequences of cDNAs encoding S-allele specific stylar RNases in self-incompatible
cultivar and its self-compatible mutant of Japanese pear, Pyrus pyrifolia Nakai. J Biochem 120:335-
345
Osaki H, Kudo A, Ohtsu Y (1998) Nucteotide sequence of seed- and pollen-transmitted double stranded RNA,
which encodes a putative RNA-dependent RNA polymerase, detected from Japanese pear. Biosci
Biotech Biochem 62:2101-2106
Palombi MA, Lombaro B, Caboni E (2007) In vitro regeneration of wild pear (Pyrus pyraster Burgsd) clones
tolerant to Fe-chlorosis and somaclonal variation analysis by RADP markers. Plant Cell Rep 26:489-
496
Perraudin G (1955) The susceptibility of fruit trees to late frosts. (in Italian). Rev Romande Agric Vitic 11:87-88
Peterson RM, Waples JR (1988) 'Gourmet' pear. HortScience 23:633
Pieniazek SA (1966) Fruit production in China. Proc 17th Int Hort Congr 4:427-456
Pierantoni L, Cho K-H, Shin I-S, Chiodini R, Tartarini S, Dondini L, Kang S-J, Sansavini S (2004)
Characterization and transferability of apple SSRs to two European pear F1 populations. Theor Appl
Genet 109:1519-1524
Predieri S, Zimmerman RH (1997) Pear mutagenesis: In vitro treatment with gamma-rays and field selection for
vegetative form traits. Euphytica 93:227-237
Predieri S, Zimmerman RH (2001) Pear mutagenesis: In vitro treatment with gamma-rays and field selection for
productivity and fruit traits. Euphytica 117:217-227
Pruss AG, Eremeev GN (1969) Drought resistance in pear varieties of diverse geographical origin (in Russian).
Trudy Prikl Bot Genet Selek 40:56-67
Pu FS, Zing XP, Xu HY, Jia IX, Fu ZC (1963) The genetic analysis of commercial characteristics of Chinese
varieties (in Chinese). Ann Scientific Rep, Res Inst Pomology, CAAS: 1-15
Pu FS, Lin SH, Chen RY, Song WQ, Li XL (1986) Studies on the karyotype of Pyrus in China. II. Acta
Horticulturae Sinica 13:87-90
Puterka GJ, Bocchetti C, Dang P, Bell RL,Scorza R (2002) Pear transformed with a lytic peptide gene for
disease control affects non-target organisms, pear psylla (Homoptera: Psyllidae). J Econ Ent 95:797-
802
Quamme HA (1984) Observations of psylla resistance among several pear cultivars and species. Fruit Var J
38:34-36
29
Quamme HA, Kappel F, Hall JW (1990) Efficacy of early selection for fire blight resistance and the analysis of
combining ability for fire blight resistance in several pear progenies. Can J Plant Sci 70:905-913
Quamme HA, Spearman GA (1983) 'Harvest Queen' and 'Harrow Delight' pear. HortScience 18:770-772
Raese JT (1994) Fruit disorders, mineral composition and tree performance influenced by rootstocks of 'Anjou'
pears. Acta Hort 367:372-379
Reimer FC (1925) Blight resistance in pears and characteristics of pear species and stocks. Oregon Agric Exper
Stat Bull 214:99
Reynoird JP, Mourgues F, Norelli J, Aldwinckle HS, Brisset MN, Chevreau E (1999) First evidence for
improved resistance to fire blight in transgenic pear epresing the attacin E gene from Hyalophora
cecropia. Plant Sci 149:13-22
Rivalta L, Dradi M, Rosati C (2002) Thirty years of pear breeding activity at Instituto Sperimentale per la
Frutticoltura of Forli: A review. Acta Hort. 596:233-238
Robbani M, Banno K, Yamaguchi K, Fujisawa N, Liu JY, Kakegawa M (2006) Selection of dwarfing rootstock
clones from Pyrus betulaefolia and P. calleryana seedlings. J Japan Soc Hort Sci 75:1-10
Roby F (1972a) Doce mutaciones en el peral Williams obtenidas por injertos de ramitas irradiadas. Rev Invest
Agropec Ser 2, 9:55-64
Roby F (1972b) Mutaciones inducida por irradiación en el peral Packham’s Triumph. In: Induced mutation and
plant improvement, International Atomic Energy Agency, Vienna. pp 475-483
Ronald WG, Temmerson HJ (1982) Tree fruits for the Prairie Provinces. Agric Can Pub 1672E
Rubtsov GA (1944) Geographical distribution of the genus Pyrus and trends and factors in its evolution. Am
Nat 78:358-366
Sansavini S (1967) Studies on cold resistance in pear varieties (in Italian). Riv Ortoflorofruttic Ital 51:407-416
Sanzol J, Herrero M (2002) Identification of self-incompatibility alleles in pear cultivars (Pyrus communis L.).
Euphytica 128:325-331
Sanzol J, Sutherland BG, Robbins TP (2006) Identification and characterization of genomic DNA sequences of
the S-ribonuclease gene associated with self-incompatibility alleles S1 to S5 in European pear. Plant
Breeding 125:513-518
Sassa H, Hirano H (1997) Nucleotide sequence of a cDNA encoding S5-RNase from Japanese pear (Pyrus
serotina Red.) Plant Physiol 113:306
Sassa H, Hirano H (1998) Style-specific and developmentally regulated accumulation of a glycosylated
thaumatin/PR5-like protein in Japanese pear (Pyrus serotina Red.). Planta 205:514-521
Sassa H, Nishuio T, Kowyama Y, Hirano H, Koba T, Ikehashi H (1996) Self incompatibility (S) alleles of the
Rosaceae encode members of a distinct class of the T2/S-ribonuclease superfamily. Mol Gen Genet
250:545-557
Sax K (1931) The origin and relationships of the pomoideae. 1. Arnold Arbor 12:3-22
Schander H (1955) On the causes of differences in weight in the seeds of pome fruits (apple and pear), I: the
relationship between seed and fruit (in German). Z Pflanzenz 34:255-306
Sekine D, Munemura I, Gao M, Mitsuhashi W, Toyomasu T, Murayama H (2006) Cloning of cDNAs encoding
cell-wall hydrolases from pear (Pyrus communis) fruit and their involvement in fruit softening and
development of melting texture. Physiol Plant 126:163-174
Serdani M, Spotts RA, Calabro JM, Postman JD, Qu AP (2006) Evaluation of USDA National clonal Pyrus
germplasm collection for resistance to Podosphaera leucotricha. HortScience 41:717-720
Shabi E, Rotem J, Loebenstein G (1973) Physiological races of Venturia pirina on pear. Phytopathology 63:41-
43
Shaw PW, Brewer LR, Wallis DR, Bus VGM, Alspach PA (2003) Susceptibility of seedling Pyrus clones to
pear sawfly (Caliroa cerasi)(Hymenoptera: Tenthredinidae) damage. N Zea J Crop Hort Sci 31:9-14
Shaw PW, Wallis DR, Alspach PA, Brewer LR, Bus VGM (2004) Pear sawfly (Caliroa cerasi) (Hymenoptera:
Tenthredinidae) host preference and larval development on six Pyrus genotypes. Zea J Crop Hort Sci
32:257-262
Shen DX, Li ZL, Zheng SQ (1979) Inheritance of fruit characteristics in pears. J Zhejiang Agric Univ 5:83-94
Shen DX, Li ZL, Zheng SQ, Chen HQ, Lin JB (1982) Studies on the correlations between juvenile period and
growth in pear seedlings (in Chinese). Acta Hort Sin 7:25-30
Shen T (1980) Pears in China. HortScience 15:13-17
Shibli RA, Ajlouni MM, Obeidat A (2000) Direct regeneration from wild pear (Pyrus syriaca) leaf explants.
Adv Hort Sci 14:12-18
Shin YU, Yim YJ, Cho HM, Yae BW, Kim MS, Kim YK (1983) Studies on the inheritance of fruit characeters
of Oriental pear, Pyrus serotina Rehder var. culta (in Korean). Res Rep Office Rural Dev 25
(Hort):108-117
Simard MH, Michelesi JC (2002) ‘Pyriam’, a new rootstock for pear. Acta Hort 596:351-355
30
Simovski K, Ristevski B, Spirovska R (1968) Effects of negative temperatures and temperature fluctuations on
pears (in Macedonian). Annu Fac Agrie Sylvie Skopjl Agric 21:125-129
Spiegel-Roy P, Alston FH (1979) Chilling and post-dormant heat requirement as selection criteria for late-
flowering pears. J Hort Sci 54:115-120
Stotz HU, Powell ALT, Damon SE, Greve LC, Bennett AB, Labavitch JM (1993) Molecular characterization of
a polygalacturonase inhibitor from Pyrus communis L. cv. Bartlett. Plant Physiol 102:133-138
Stushnoff C, Garley B (1982) Breeding for cold hardiness. p. 189-199. In: van de Zwet T, Childers NF (eds.)
The pear. Horticultural Publ, Gainesville, FL
Subhadrabandhu S (1999) Deciduous fruit production in Thailand. In: Papademetriou MK, Herath EM (eds.)
Deciduous Fruit Production in Asia and the Pacific. RAP Publication: 1999/10. (Bangkok, Thailand:
Food and Agriculture Organization of the United Nations Regional Office For Asia and the Pacific)
Suzuki Y, Maeshima M, Yamaki S (1999) Molecular cloning of vaculor H(+)-pyrophosphatase and its
expression during the development of pear fruit. Plant Cell Physiol 40:900-904
Takasaki T, Moriya Y, Okada K, Yamamoto K, Iwanami H, Besso H, Nakanishi T (2006) cDNA cloning of
nine S alleles and establishment of a PCR-RFLP system for genotyping European pear cultivars. Theor
Appl Genet 112:1543-1552
Tateishi A, Inoue H, Shiba H, Yamaki S (2001) Molecular cloning of β-galactosidase from Japanese pear (Pyrus
pyrifolia) and its gene expression with fruit ripening. Plant Cell Physiol 42:492-493
Tateishi A, Mori H, Watari J, Nagashima K, Yamaki S, Inoue H (2005) Isolation, characterization, and cloning
of α-L-Arabinofuranosidase expressed during fruit ripening of Japanese pear. Plant Physiol 138:1653-
1664
Terakami S, Shoda M, Adachi Y, Gonai T, Kasumi M, Sawamura Y, Iketani H, Kotobuki K, Patocchi A,
Gessler C, Hayashi T, Yamamoto T (2006) Genetic mapping of the pear scab resistance gene Vnk of
Japanese pear cultivar Kinchaku. Theor Appl Genet 113:743-752
Thibault B, Hermann L, Belouin A, Mangin B (1988) Inheritance of some agronomi cal traits in pear. Acta Hort
224:199-209
Thibault B, Lecomte P, Hermann L, Belouin A (1987) Assessment of the susceptibility of Erwinia amylovora of
90 varieties or selections of pear. Acta Hort 217:305-309
Thompson JM, van der Zwet T, Ditto WA (1974) Inheritance of grit content in fruits of Pyrus communis L. J
Amer Soc Hort Sci 99:141-143
Thompson JM, Zimmerman RH, van der Zwet T (1975) Inheritance of fire blight resistance in pear, I: a
dominant gene, Se, causing sensitivity. J Hered 66:259-264
Tufts WP, Day LH (1934) Nematode resistance of certain deciduous fruit tree seedlings. Proc Am Soc Hort Sci
31:75-82
Tukey HB (1964) Dwarfed fruit trees. Macmillan, NY
Tuz AS (1972) The inheritance of the dwarf growth factor in pear, Pyrus domestica Medii (in Russian).
Genetika 8:16-20
Ushijima K, Sassa H, Hirano H (1998) Characterization of the flanking regions of the S-RNase genes of
Japanese pear (Pyrus serotina) and apple (Malus x domestica). Gene 211:159-167
van der Zwet T, Bell RL (1990) Fire blight susceptibility in Pyrus germplasm from Eastern Europe. HortScience
25:566-568
van der Zwet T, Keil HL (1979) Fire blight: a bacterial disease of rosaceous plants. Agric Handb 510 USDA,
Washington, DC
van der Zwet T, Oitto WA, Brooks HJ (1970) Scoring system for rating the severity of fire blight in pear. Plant
Dis Rep 54:835-839
Vavilov NI (1951) The origin, variation, immunity and breeding of cultivated plants. Translated by K Start
Chron Bot 13:1-366
Vavra M, Orel V (1971) Hybridization of pear varieties by Gregor Mendel. Euphytica 20:60-67
Verhaegh JJ, Visser T, Kellerhals M (1988) Juvenile period of apple seedlings as affected by rootstock, bud
origin and growth factors. Acta Hort 224:133-139
Visser T (1951) Floral biology and crossing technique in apples and pears (in Dutch). Meded Dir Tuinb 14:707-
726
Visser T (1967) Juvenile period and precocity of apple and pear seedlings. Euphytica 16:319-320
Visser T (1976) A comparison of apple and pear seedlings with reference to the juvenile period, II: mode of
inheritance. Euphytica 16:339- 342
Visser T, De Vries DP (1970) Precocity and productivity of propagated apple and pear seedlings as dependent
on the juvenile period. Euphytica 19:141-144
Visser T, Schaap AA, De Vries DP (1968) Acidity and sweetness in apple and pear. Euphytica 17:153-167
31
Visser T, Verhaegh JJ, De Vries DP (1971) Pre-selection of compact mutants induced by x-ray treatment of
apple and pear. Euphytica 20:195-207
Visser T, Verhaegh JJ, De Vries DP (1976) A comparison of apple and pear seedlings with reference to the
juvenile period, I: seedling growth and yield. Euphytica 25:343-351
Vondracek J (1982) Pear cultivars resistant to scab. p. 420-424. In: van der Zwet T, Childers NF (eds.) The pear.
Horticultural Publ, Gainesville, FL
Wang YL (1990) Pear breeding in China. Plant Breeding Abstr 60:877-879
Wang YL, Wei WD (1987) Studies on the inheritance of commercial characteristics in pear crossed seedlings
(in Chinese). 1. Decid Fruits 2:1-4
Webster AD (1998) A brief review of pear rootstock development. Acta Hort 475:135-141
Wellington R (1913) Inheritance of the russet skin in the pear. Science 37:156
Westigard PH, Westwood MN, Lombard PB (1970) Host preference and resistance of Pyrus species to the pear
psylla, Psylla pyricola Foerster. J Amer Soc Hort Sci 95:34-36
Westwood MN (1976) Inheritance of pear decline resistance. Fruit Var J 30:63-64
Westwood MN (1982) Pear germplasm of the new national clonal repository: Its evaluation and uses. Acta Hort
124:57-65
Westwood MN, Bjornstad HO (1971) Some fruit characteristics of interspecifi hybrids and extent of self-
sterility in Pyrus. Bull Torrey Bot Club 98:22-24
Westwood MN, Westigard PH (1969) Degree of resistance among pear species to the woolly pear aphid,
Eriosoma pyricola. J Amer Soc Hort Sci 94:91-93
White AG, Alspach PA (1996) Variation in fruit shape in three pear hybrid progenies. N Zeal J Crop Hort Sci
24:409-413
White AG, Brewer LR (2002) The New Zealand pear breeding project. Acta Hort 596:239-242
White AG, Alspach PA, Weskett H, Brewer LR (2000) Heritability of fruit shape in pears. Euphytica 112:1-7
Whitesides SK, Spotts RA (1991) Susceptibility of pear cultivars to blossom blast caused by Pseudomonas
syringae. HortScience 26:880-882
Wu H-Q, Zhang S-L, Qu H-Y (2007) Molecular and genetic analysis of S4
SM RNase allele in Japanese pear
‘Osa-Nijisseiki’ (Pyrus pyrifolia Nakai). Plant Breeding 126:77-82
Yamada K, Kojima T, Bantog N, Shimoda T, Mori H, Shiratake K, Yamaki S (2007) Cloning of two isoforms
of soluble acid invertase of Japanese pear and their expression during fruit development. J Plant Phsiol
164:746-755
Yamamoto T, Kimura T, Sawamura Y, Kotobuki K, Ban Y, Hayashi T, Matsuta N (2001) SSRs isolated from
apple can identify polymorphism and genetic diversity in pear. Theor Appl Genet 102:865-870
Yamamoto T, Kimura T, Sawamura Y, Manabe T, Kotobuki K, Hayashi T, Ban Y, Matsuta N (2002a) Simple
sequence repeats for genetic analysis in pear. Euphytica 124:129-137
Yamamoto T, Kimura T, Shoda M, Ban Y, Hayashi T, Matsura N (2002b) Development of microsatellite
markers in the Japanese pear (Pyrus purifolia Nakai). Mol Biol Notes 2:14-16
Yamamoto T, Kimura T, Shoda M, Imai T, Saito T, Sawamura Y, Kotobuki K, Hayashi T, Matsuta N (2002c)
Genetic linkage maps constructed by using an interspecific cross between Japanese and European
pears. Theo Appl Genet 106:9-18
Yancheva SD, Shlizerman LA, Golubowicz S, Yabloviz Z, Perl A, Hanania U, Flaishman MA (2006) The use
of green florescent protein (GFP) improves Agrobacterium-mediated transformation of ‘Spadona’ pear
(P. communis L.). Plant Cell Rept 25:183-189
Yu DY, Zhang P (1979) Sinkiang pears, a new series of cultivars of pears in China. Acta Hort 6:27-32
Zavoronkov PA (1960) Breeding winter-hardy pear varieties (in Russian). Sadovodstvo horticulturae 11:28-31
Zeven AC, Zhukovsky M (1975) Dictionary of cultivated plants and their centres of diversity. Centre for
Agricultural Publishing and Documentation, Wageningen
Zhejiang Agricultural University (1977) Inheritance of some characters in pear seedlings. Proc Nat Acad Conf
Pears 1976:114-121
Zhejiang Agricultural University (1978) Studies on the inheritance of the precocity in pears (in Chinese). Acta
Genetica Sinica 5:220-226
Zhu LH, Ahlman A, Welander W (2003) The rooting ability of the dwarfing rootstock BP10030 (Pyrus
communis) was significantly increased by introduction of the rolB gene. Plant Sci 135:829-835
Zhu LH, Welander W (2000) Adventitious shoot regeneration of to dwarfing pear rootstocks and the
development of a transformation protocol. J Hort Sci Biotech 75:745-752
Zielinski QB (1963) Precocious flowering of pear seedlings carrying the Cardinal Red color gene. J Hered
54:75-78
Zielinski QB, Reimer FC, Quackenbush VL (1965) Breeding behavior of frui characteristics in pears, Pyrus
communis L. Proc Amer Soc Hort Sci 86:81-87
32
Zielinski QB, Thompson MM (1967) Speciation in Pyrus: Chromosome number and meiotic behavior. Bot
Gazette 128:109-112
Zimmerman RH (1972) Juvenility and flowering in woody plants: a review. HortScience 7:447-455
Zimmerman RH (1976) Transmittance of juvenile period in pears. Acta Hort 56:219-224
Zimmerman RH (1977) Relation of pear seedling size to length of the juvenile period. J Amer Soc Hort Sci
102:443-447
... The positive properties of grafting pears on quince rootstocks are controlling the growth of trees, precocity, and increase in yield and fruit quality. However, quince clonal rootstocks have a low tolerance to chlorosis, fire blight (caused by the Erwinia amylovora bacterium), and calcareous soils, and are not being firmly attached to the soil as pear rootstocks [20]. Nowadays, improved pear orchards are based on a high density planting system (HDP) around 2000 -5000 trees in a hectare. ...
Article
Full-text available
Abstract The present study was carried out to determine the effect of quince rootstocks (BA29, QA and MC) on plant growth, fruit quality and yield of ‘Santa Maria’ pear cultivar in 2020 and 2021. In this study, significant differences were determined by all growth, yield and quality parameters except for trunk diameter (TD), trunk cross-sectional area (TCSA), canopy volume, fruit stalk thickness, acidity, total soluble solids (TSS) in terms of research years and rootstocks from analysis of variance. Tree height, annual shoot length, leaf area, fruit weight, fruit width, and fruit length was lower in the MC rootstock than in the others. In the case of research years, out of annual shoot growth in 2020, all other parameters were higher in 2021. Fruit stalk length, fruit stalk thickness and fruit firmness were higher on MC rootstock than BA29 and QA. There was an insignificant difference between rootstocks in terms of TSS. The highest L* was in QA and MC; chroma was in QA and hue angle was in BA29 rootstock. The number of fruit was the highest (139.01 pieces/tree) in BA29 in 2021 and the lowest in MC (11.97 pieces/tree) in 2020. The highest yield per tree (30.62 kg), yield per hectare (58484.2 kg), yield per trunk cross-sectional area (2.24 kg/cm2) and yield per canopy volume (71.21 kg/m3) were measured in QA in 2021. In conclusion, the yield efficiency of ‘Santa Maria’ pear cultivar under high density planting on QA was better than other combinations in 2021. Since the plants on which the research is carried out are still young, it may be recommended to continue the research for many years to determine the most suitable rootstock.
Article
Full-text available
In the Mediterranean basin, agriculture and other forms of human land use have shaped the environment since ancient times. Intensive and extensive agricultural systems managed with a few cultured plant populations of improved varieties are a widespread reality in many Mediterranean countries. Despite this, historical cultural landscapes still exist in interior and less intensively managed rural areas. There, ancient fruit tree varieties have survived modern cultivation systems, preserving a unique genetic heritage. In this study, we mapped and characterized 106 living fruit trees of ancient varieties in the Basilicata region of southern Italy. Tree ages were determined through tree ring measurements and radiocarbon analyses. We uncovered some of the oldest scientifically dated fruit trees in the world. The oldest fruit species were olive (max age 680 ± 57 years), mulberry (647 ± 66 years), chestnut (636 ± 66 years), and pear (467 ± 89 years). These patriarchs hold a unique genetic resource; their preservation and genetic maintenance through agamic propagation are now promoted by the Lucan Agency for the Development and Innovation in Agriculture (ALSIA). Each tree also represents a hub for biodiversity conservation in agrarian ecosystems: their large architecture and time persistence guarantee ecological niches and micro-habitats suitable for flora and fauna species of conservation significance.
Article
Full-text available
Pyrus communis L. is an important temperate fruit with high nutritional and economic value. Italy, as the largest pear producer in the EU and second in the world, has a particularly rich germplasm characterized by hundreds of local varieties. The Lazio Region was the first Italian region to start programs focused on safeguarding varieties at risk of extinction and has started a massive census of fruit varieties by combining molecular data and productive-territorial information. In this study, 311 pear accessions collected from the five provinces of the Lazio region were genetically characterized by the means of nine simple sequence repeat (SSR) markers, eight of which were recommended by the European Cooperative Programme for Plant Genetic Resources (ECPGR). The SSR analysis revealed 250 unique genotypes and 14 cases of synonymies, namely, accessions with different names but identical molecular profiles (100% genetic similarity). The microsatellite set showed a high polymorphism information content (PIC; mean PIC=0.77) and an exceptionally high discrimination power (DP = 0.99), making it particularly efficient both for the study of genetic diversity and for the prediction of the degree of ploidy. Notably, 20% of the accessions displayed triallelic profiles (i.e., accessions having ≥2 loci with a third allele), while the remaining accessions were diploids. These results were further confirmed by flow cytometry data analysis. Standardization of the molecular analyses at the national and international levels and harmonization of the SSR sets used for germplasm characterization are of paramount importance for producing molecular profiles useful for registration and variety maintenance.
Article
Full-text available
The total diversity of bacterial and fungal communities associated with the phyllosphere (fruits and leaves) of the ‘Williams’ pear variety was analyzed in two phenological stages during fruit development and maturation. The antagonistic potential of autochthonous bacterial and yeast isolates against phytopathogenic fungi was also evaluated. A metabarcoding approach revealed Pantoea, Sphingomonas, Hymenobacter, Massilia, and Pseudomonas as dominant bacterial constituents of the pear phyllosphere, whilst most abundant among the fungal representatives identified were Metschnikowia, Filobasidium, Aureobasidium pullulans, Botrytis cinerea, and Taphrina. The traditional culturable approach revealed that the Pseudomonas genus with P. graminis, P. putida, and P. conge-lans was most prevalent. The most frequently cultivated fungal representatives belonged to the genus Fusarium with six identified species. A broad range of the antagonistic activity was detected for the Hannaella luteola and Metschnikowia pulcherrima yeasts, significantly affecting the growth of many fungal isolates in the range of 53–70%. Fusarium sporotrichioides was the most susceptible fungal isolate. The autochthonous antagonistic yeasts H. luteola and M. pulcherrima might be powerful biological control agents of postharvest diseases caused by Fusarium spp. and common pathogens like Monilinia laxa, Botrytis cinerea, Alternaria tenuissima, and Cladosporium cladosporioides.
Article
Full-text available
To evaluate the phylogenetic relationships among subtribe Pyrinae of Rosaceae, 50 different genotypes and cultivars of pome fruit trees were collected from various locations in Iran, and analysed using random amplified polymorphic DNA (RAPD) markers. Altogether, 85 polymorphic fragments were produced by 11 random 10-mer primers. The number of produced bands varied from 4 to 12 for each primer, 7.72 on average. The lowest Jaccard's genetic similarity coefficient was scored between apple cv. ‘Akan’ and a wild Pyrus syriaca pear (0.04), and the highest similarity was observed between two genotypes of P. syriaca (0.97). Cluster analysis using the unweighted pair group method with arithmetic mean (UPGMA) properly separated the accessions and divided them into three major groups, including Malus, Pyrus, Mespilus and Crataegus. Members of Mespilus and Crataegus were further separated and formed two subgroups. Analysis of the genetic structure, using STRUCTURE and phylogenetic relationship using TreeView software confirmed the results of the UPGMA dendrogram and properly distinguished different genera. Among different genera studied in our experiment, Malus was the most distinct one and showed high genetic distance from other genera. Generally, the results from the present study provide valuable information about the genetic relationship among important pome fruit trees in Iran and indicate that, although Mespilus is genetically close to the Crataegus genus, it has relatively high similarity with Pyrus and Malus. Moreover, outcomes of this research confirmed that RAPD molecular markers are an appropriate means for evaluation of the genetic relationship among various genera of pome fruit trees.
Article
Full-text available
The European pear psylla, Cacopsylla pyri L. (Hemiptera: Psyllidae), is one of the most serious arthropod pests of pear. Since proper control of this pest is essential, better understanding of the complex plant-pest relationship is mandatory. This research deals with constitutive polyphenolic profiles in leaves of 22 pear cultivars of diverse origin (P. communis, P. pyrifolia and P. pyrifolia x P. communis) and different resistance to psylla. The study was designed to show which differences in the polyphenolic profile of leaves from resistant and susceptible pear cultivars could be utilized as information in subsequent breeding programs. The results demonstrated that the leaves of Oriental pear cultivars contained much higher amounts of p-hydroxybenzoic acid, ferulic acid, aesculin and naringin, that, together with detected 3-O-(6"-O-p-coumaroyl)-hexoside, apigenin, apigenin 7-O-rutinoside and hispidulin, indicated a clear difference between the species and might represent phenolics responsible for psylla resistance.
Article
The Northwest of Iran has a high diversity of endemic Pyrus species. In this study, considerable variability in morphological and chemical characteristics was observed in the fruit of 19 local Pyranshahr pear accessions. Fruit samples of three accessions (14, 7 and 10) had higher contents of total soluble solids (TSS > 23 %) and total sugar (SC ~11.5 %) than other accessions with a mean SC of 9 %. There was significant positive correlation between SC and TSS (r = 0.40) as well as between SC and pH (r = 0.46). Cluster analysis of the studied parameters, which were used to ascertain the level of diversity, differentiated accessions into seven clusters with significant variability among them. This research shows that pear genetic resources in the Northwest of Iran show rich genetic diversity and could be useful germplasm for commercial exploitation, for use in breeding programs to improve existing cultivars, or for developing new cultivars.
Article
Full-text available
Fire blight caused by the bacterium Erwinia amylovora, is one of the most catastrophic diseases of plants in the Rosaceae subfamily Pomoideae. The objective of this study was to determine the resistance level of 30 pear cultivars with mixing of four Erwinia amylovora strains. Disease development was assessed according to severity indices (I.V.S.) based on the progression of necrosis rate through shoots, beginning at the inoculation point at 3, 6, 9, 12 and 18 days after the inoculation. Analysis of variance showed significant differences among evaluated cultivars, indicating a wide range of resistance to the fire blight pathogen. Pear cultivars showed considerable variations in the resistance to fire blight with the severity of infection ranging from 0.4% to 100%. Cultivars in this study were divided into four groups including very resistant (0-10%), semi-resistant (11%-50%), susceptible (51%- 80%) and very susceptible (81%- 100%) according to the percentage of the length of necrosis to the total shoot length. Three cultivars including ‘Dare Gazi’, ‘Harrow Sweet’ and ‘Natanzi’ were identified as very resistant, 14 cultivars as semi-resistant, 10 cultivars as susceptible and another three cultivars including ‘Mohamad Ali’, ‘Duchesse’ and ‘KS6’ were recognized as very susceptible. Results obtained from specific primers link to fire blight showed that RLG1 and CH02c02b primers produced resistance allele in Dargazi and some of other cultivars.
Article
Full-text available
Species of the genus Pyrus, commonly referred to as pears, are pome fruits related to apples and quince. There are at least 26 widely recognized primary species and 10 naturally occurring interspecific hybrid taxa, which are distributed in Europe, temperate Asia, and mountainous areas of northern Africa. There is little well-documented information on the conservation status of pear species, although several taxa are listed in the IUCN Red List. Conservation efforts tend to be mostly in ex situ collections of cultivars, and more seed collection from wild populations is desirable. All species of Pyrus are intercrossable and there are no major incompatibility barriers to . There are four main cultivated species, which have been the focus of most intra- and interspecific hybridization for the development of scion cultivars. However, several other species are sources of valuable traits, principally for the development of rootstocks, or directly for ornamental cultivars. Intergeneric hybridization is difficult and hybrid lethality and low fertility of the hybrids is common. However, ×Sorbopyrus auricularis is a naturally occurring intergeneric hybrid of mountain ash and pear, and ×Pyronia veitchii is a hybrid of Pyrus pyrifolia and Cydonia oblonga. Various apple × pear hybrids have also been produced, especially when special techniques such as embryo rescue are used.
Article
Full-text available
Fruit quality and flavour are important targets in all pear breeding programmes. Perceived sweetness is directly influenced by the amount and type of sugar accumulated in fruit. Limited information is available on sugar composition in pear fruit and published studies have been completed using cultivars rather than breeding populations. The objective of this research was to determine the quantitative genetic parameters of sugar content in fruit of interspecific hybrids from families making up a pear breeding population. Glucose, fructose, sucrose and sorbitol contents were measured in mature fruit. Most of the sugars, except for sorbitol, showed genetic variability and a relatively high (i.e., > 0.5) ratio between the estimated additive genetic variance and the total variance. Sorbitol showed a high negative genetic correlation (–0.65) with fructose. It could be suggested that the main product of sorbitol conversion was fructose. Sucrose showed a negative genetic correlation with glucose (–0.37) and fructose (–0.16), which would be expected given that sucrose is metabolised into fructose and glucose. Two parents with 100 % European parentage showed the highest empirical breeding values (eBV)s for fructose and total sugars. The parent with 100 % Asian parentage showed the lowest eBV for sorbitol. The mean percentages of the sugars across the entire population were: glucose 13 %, fructose 59 %, sucrose 8 % and sorbitol 20 %, indicating fructose was the main sugar with sucrose less prominent.
Article
Full-text available
‘Harvest Queen’ and ‘Harrow Delight’ are high-quality pear cultivars introduced for early fresh-market and home garden use. At Harrow, ‘Harvest Queen’ ripens during the 3rd week of August, one week before ‘Bartlett’, and ‘Harrow Delight’ ripens during the 2nd week of August, 2 weeks before ‘Bartlett’ (Table 1). Both cultivars are resistant to fire blight caused by Erwinia amylovora (Burr.) Winsl. et al., and are adapted to conditions in southern Ontario.
Article
Full-text available
The age at which flowering began and the stem diameter at various ages were recorded for more than 9,000 pear seedlings planted in orchards at Beltsville. The juvenile period for individual seedlings varied from 2 to 10 years, with a number of seedlings still not flowering after 8 to 10 years. An overall negative correlation of stem diameter with length of the juvenile period was found. Within planting years, this relation was significant only in certain years; within progenies, the relation was significant less than half the time; within a specific cross repeated in different years, the relation varied considerably from year to year. Stem diameter can be successfully used as a preselection criterion for early flowering (short juvenile period) only when these two characteristics are significantly correlated. Since this condition does not exist in most crosses under the growing conditions at Beltsville, stem diameter is not a valid predictor of early flowering.
Article
Full-text available
No consistent trend towards increased inbreeding has existed within the U.S. Department of Agriculture pear breeding program over 17 years of crosses based on the mean inbreeding coefficient and the percentage of non-inbred progenies. Selections did not tend to be more or less inbred than the population of all seedings. There was consistent, but small, trend towards a reduction in seedling vigor with increased levels of inbreeding as measured by 5th year stem diameter. A significant positive, but small, association between increased inbreeding and improved flavor, grit, and texture were observed, even after correction for the effects of parental values for these characters. Limited inbreeding does not adversely affect improvement of fruit quality and appears to be of some benefit in facilitating selection of favorable alleles.
Article
China has been the world's largest producer of eggs for the last 30 years. There have been considerable recent changes in the structure of the egg industry due to rapid economic growth, improved supply chains and favourable prices. Since 2012, revenue from poultry has increased by 8.7% on average. Annual growth rate of egg production has been approximately 0.6 million tonnes per year from 2000 to 2016 (FAO, 2017), and 2016 saw the total eggs produced reaching a peak of 31 million tonnes. Egg production in China continues to provide the population with a significant proportion of their intake of high quality dietary protein. Egg consumption is expected to continue to increase with the rise in the urban population. It is probable that the highest increase in demand for eggs will be for ‘out of home’ consumption and in processed food products. These levels of production are associated with the spread of intensive systems. The poultry sector is no longer dominated by hundreds of millions of smallholders keeping birds as a side-line activity and many small farmers have ceased production. Chinese consumers are becoming more focused on the quality and safety of eggs. Future developments in the egg production are expected to concentrate on quality, safety and traceability of eggs.