{"id":9264,"date":"2023-09-25T05:13:50","date_gmt":"2023-09-25T03:13:50","guid":{"rendered":"https:\/\/marijuanagrowing.com\/?p=9264"},"modified":"2024-01-31T13:40:29","modified_gmt":"2024-01-31T12:40:29","slug":"breeding-chapter-25","status":"publish","type":"post","link":"https:\/\/marijuanagrowing.com\/breeding-chapter-25\/","title":{"rendered":"Breeding – Chapter 25"},"content":{"rendered":"\n

Casual and serious cannabis breeders working with Mother Nature use selective breeding have transformed a wild plant into countless varieties of medicinal cannabis. New varieties are easy, fun, and satisfying to create. Home cannabis gardeners breed plants for many reasons\u2014to increase yield, cannabinoid profile, disease and pest resistance, and so forth. Most often, modern medicinal cannabis has been bred for high cannabinoid content (specifically THC), for heavy yield, for early harvest indoors, and occasionally for harsh outdoor climates. Today, more attention is being given to other cannabinoids, such as CBD, and cultivation stress\u2014resistance to diseases and pests and tolerance of drought and cold.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Nevil, founder of the Seed Bank of Holland, traveled the world to find the best cannabis seeds.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Countless cannabis gardeners are patiently applying their imagination and creativity using basic breeding (sexual propagation) techniques to create new generations of seed. Breeding is easy, low-cost, fun, and exciting. It requires minimal time and skill but results in big rewards.<\/p>\n\n\n\n

The basics outlined in this chapter show any gardener how to start breeding and create new generations of viable medicinal cannabis seeds. Indoor, backyard, and greenhouse breeding is essential to acclimate varieties to local climates and to increase resistance to diseases, pests, and stress. Indoor and greenhouse cultivation can expedite outdoor breeding programs. Unfortunately, medical cannabis seeds are not available in some jurisdictions, and indoor breeding is a matter of necessity for patients and caregivers.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This colorful \u2018Pakistan Chitral Kush\u2019 from CannaBioGen is a pure landrace line from Chitral, Pakistan. It is also on the cover of this book!<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Seated in the center in white, Sensi Seeds owner Ben Dronkers is pictured collecting seeds in Afghanistan.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"cannabis<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The Internet has changed the availability of genetics (seeds) worldwide.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

On the surface, breeding is simple: pollen from a male plant is used to fertilize (pollinate) a female plant, and seeds result. In fact, this is what most cannabis crosses are based upon\u2014crossing a cannabinoid-potent male with a cannabinoid-potent female plant, often referred to as \u201cpollen chucking.\u201d However, a few important genetic details are involved.<\/p>\n\n\n\n

Crossing cannabis plants can become very complex, and plants take time to grow to express their genes. Modern plant breeders study statistics, chemistry, plant physiology, microbiology, and other disciplines to understand and breed plants. The basic information presented here is in an easy-to-understand format so it is as straightforward to reference as possible. Cannabis breeding is a long-term process, and the possible outcomes of progeny are infinite.<\/p>\n\n\n\n

<\/div>\n\n\n\n
\"\"<\/figure>\n\n\n\n
<\/div>\n\n\n\n

To date, much of the cannabis breeding in the USA has been done by clandestine breeders such as Mel Frank, who introduced varieties such as \u2018Afghani\u2019, \u2018Nepalese\u2019, and \u2018Mexican\u2019, as well as many Colombian varieties.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Much more scientific information about plant breeding is available on the Internet, including modern genetic principles by Gregor Mendel and inspiring stories from American breeders such as Luther Burbank. Be sure to read about Dr. Kevin McKernan, who mapped the cannabis genome!<\/p>\n\n\n\n

Please continue studying after you have digested this basic chapter. Remember, breeding cannabis is a long-term endeavor\u2014even in an indoor garden room where four generations are possible.<\/p>\n\n\n\n

The cannabis world today consists of millions of small gardeners in more than 200 countries around the world*. Many gardens exist in unstable and difficult growing conditions. Adopting new cannabis varieties developed for cultivation in faraway climates and indoor grow rooms does not meet the needs of many outdoor gardeners and their local climates. The Internet* and cannabis gardeners\u2019 ingenuity have broken ground for gardeners everywhere to participate in the breeding process. Breeding goals can now be identified and defined by local gardeners instead of large seed companies or fad-driven advertising.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The \u2018Afghani\u2019 \u00d7 \u2018Congolese\u2019 plants in the foreground are the product of selective breeding. The miscellaneous sativa plants in the background represent wild cannabis plants. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Choosing the proper male breeding stock is half of the equation. Choose the best males possible that exhibit the desirable traits you want.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

For example, the following YouTube video has received more than 5 million hits from more than 180 countries: www.youtube.com\/user\/jorgecervantesmj<\/a>.<\/p>\n\n\n\n

Most gardeners prefer to acquire seeds from a reputable seed company, and too often assume these seeds are consistently of the highest genetic quality. Often, commercially available seeds are produced by a few large seed makers who wholesale to resellers, many of them with Internet sites and retail store outlets. Other seeds are produced by basement growers, small startups, or small- to medium-sized established companies.<\/p>\n\n\n\n

Few small companies can maintain a library of true-breeding stock and breed stable parent stock over time. Stock produced by large companies usually fills demand and maintains standards.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This beautiful garden of \u2018Sonoma Coma\u2019 females is the result of years of selective breeding in a valley full of vineyards in California.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Individual male flowers that appear on a female flower bud are often called “bananas” because they look like little bananas.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This ‘African’ grown in 1977 has a small male flower on this predominantly female flower bud. This is an intersex male flower.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Unforgettable Qualities of Cannabis<\/h3>\n\n\n\n

Cannabis has a unique set of qualities that influence its life cycle and reproduction. Working with these natural qualities is essential to breeding cannabis. Please remember them when working on all breeding projects.<\/p>\n\n\n\n

Cannabis is photoperiodic reactive<\/strong>, flowering under 12 hours of uninterrupted darkness and 12 hours of light, which gives breeders the ability to grow up to 6 crops every year. These characteristics facilitate fast-track breeding because 6 years of crosses can be completed in 1 year.<\/p>\n\n\n\n

Cannabis is dioecious<\/strong>; it produces each of the male (staminate; stamen-bearing) and female (pistillate; pistil-bearing) sexual organs on different individuals, and it is one of the few annual dioecious plants. This quality makes it very easy to cross an individual male cannabis plant with an individual or population of females.<\/p>\n\n\n\n

Male (staminate) and female (pistillate) plants are easy to distinguish. See close-up details of male and female flowers along with full descriptions below in this chapter.<\/p>\n\n\n\n

Monoecious<\/strong> varieties produce both staminate and pistillate flowers on the same plant. Monoecious varieties are mainly used for hemp seed production. Monoecious varieties do not make good medical cannabis breeding stock. Plants with both male and female flowers are often called by the misnomer “hermaphrodite.”<\/p>\n\n\n\n

Outcrossing plant: <\/strong>Outcrossing cannabis grows best when plants cross with other plants with different genes. Corn, dogs, cannabis, and people are all outcrossers. Backcrossing and inbreeding cannabis for too many generations will be detrimental.<\/p>\n\n\n\n

Intersexuality<\/strong> occurs when flowers of a male plant grow on a predominantly female plant, or when female flowers grow on a predominantly male plant. It is a trait that can be caused by both genetic and environmental factors. Intersex plants with the inherited gene grow flowers from both sexes on the same plant even in perfect growing conditions. Intersex plants are often called by the misnomer “hermaphrodite.”<\/p>\n\n\n\n

Indoors, plants are easily exposed to stress\u2014inconsistent or extreme temperature, light cycles, fertilizer, pH, etc.\u2014and this stress can cause female plants to grow an occasional male flower. Breeders prefer not to perpetuate the intersex gene, and when possible, they eliminate it. A single male flower on a female plant can pollinate a big part of the female. See “Feminized Seeds,” on page 521.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Physiology of Male and Female Flowers<\/h2>\n\n\n\n

Male Flowers<\/h3>\n\n\n\n

Male (staminate) cannabis flowers are about 0.25 to 0.5 inches (0.6\u20131.3 cm) long, and thousands of individual flowers can develop on a large plant. Most of the flowers develop in loose clusters (cymes or cymose panicles) of about 5 or 10 flowers each, borne on tiny branches and their side (lateral) branches. The clusters can pile atop each other to form dense aggregates of hundreds of individual flowers, particularly at the ends of stems and branches.<\/p>\n\n\n\n

Each male flower’s calyx consists of 5 tepals (sometimes identified as “sepals”)\u2014usually white, yellowish, or greenish, but often tinged purple\u2014that might be described as “petals,” and 5 pendulous stamens that bear pollen in sacks called anthers. Anthers hang by a thin, threadlike filament, and together, filament and anther make up the stamen. Once mature, 2 openings on opposite sides of each anther open zipper-like, starting at their base, to slowly release their pollen into the wind, carrying it (hopefully) to stigmas. It has been estimated that the thousands of flowers on a single male can release more than 500 million pollen grains.<\/p>\n\n\n\n

Unopened male flower clusters remind some growers of tiny grape clusters, and fresh anthers look somewhat like bunches of tiny bananas. Male flowers are simply called male flowers or male flower clusters, and the pollen holders are referred to as either stamens or anthers.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This drawing shows the main parts of a male cannabis plant.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

‘Jack Herer’ male (staminate) flowers fully formed in clusters but not yet open. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Outer tepals on the ‘Skunk #1’ staminate calyxes have separated, exposing the pollen-containing anthers. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

As anthers on this ‘Jack Herer’ mature, they rupture to disperse superfine grains of pollen into the air. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Female Flowers<\/h3>\n\n\n\n

Each female marijuana flower has 2 stigmas that protrude from the single ovule enclosed in the bracts; fresh stigmas are “fuzzy” (hirsute), about 0.25 to 0.5 inches (0.6\u20131.3 cm) long, and usually white, but sometimes they are yellowish or pink to red, and, rarely, lavender to purple. Stigmas (stigmata is another botanical plural) are the pollen catchers. Stigmas are often misidentified as pistils. By definition, a pistil is all of the reproductive parts of a flower\u20142 stigmas and an ovule make up the female pistil. Each flower then has only 1 pistil but has 2 stigmas. The term is misused in much of popular culture, which describes a single cannabis flower as having 2 pistils.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Staminate anthers on this ‘Skunk #1’ grown in 1987 continue to split open and disperse more and more pollen into the air. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Anthers hang in the wind after dispersing all their pollen throughout several days. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Pollen from this ‘Skunk #1’ has been dispersed from anthers in the foreground. Staminate calyxes in the background will disperse pollen in the near future. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Compare this tiny male flower to the size of a US penny. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Stigmas begin dying after pollination and begin to turn rust-colored about 3 days later; if not pollinated, as with sinsemilla (seedless buds), stigmas begin to die when they are about 4 or 5 weeks old. Upon landing on a stigma, a pollen grain germinates and begins to grow a pollen tube through the style passageway to pass its DNA to join the ovule\u2019s DNA (the 2 stigmas of each female flower make up a bifid style). The fertilized ovule becomes a fruit, essentially a single seed (an achene). The perianth, which includes the calyx, tightly clasps the seed and often contains tannins, which give mature seeds their mottled or spotted coat. Between a thumb and finger you can rub the perianth off of seeds. A well-pollinated single bud develops dozens of seeds, a cola easily holds many hundreds, and even a small, but thoroughly pollinated female can bear thousands of seeds.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This drawing shows the main parts of a female cannabis plant.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Each female flower has a single ovule partially enclosed in its perianth, which is encapsulated by bracteoles, which are covered by a whorl of bracts. The bracts and bracteoles are small, modified leaves that enclose and protect the seed in what some growers refer to as the seedpod.<\/p>\n\n\n\n

Bracts contain the highest concentration of THC and other cannabinoids of any plant part and about 50 percent of a plant\u2019s total THC. The perianth and its calyx contain no THC.<\/p>\n\n\n\n

By definition, a perianth consists of a corolla and a calyx. In more familiar showy flowers, the corolla is the brightly colored petals we generally appreciate when looking at flowers, and the calyx is the smaller green cup (sepals) at the flower\u2019s base. Bright showy colors, large flower sizes, and enticing fragrances have naturally evolved to attract insects such as bees and flies, or animals such as birds and bats that collect and transfer pollen (unintentionally) to other flowers. Cannabis flowers are not brightly colored, large, or enticingly fragrant (at least to most nonhumans); cannabis plants are wind-pollinated with no need to attract insects or animals to carry males\u2019 pollen to female flowers; cannabis plant parts never naturally evolved into colorful, attractive, or showy parts. Cannabis breeders, though, do breed for fragrance and color once cannabinoid content is firmly established.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The seed bract still covers the perianth, pistillate calyx, gametes, and ovule connected to a pair of stigmas. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The seed bract has been stripped away to reveal the perianth and pistillate calyx, both of which appear transparent, covering the gametes and ovule. Note the white stigmas have not been pollinated. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Stigmas start dying back as soon as pollen is ushered down the shaft to unite with the ovule below. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

The cannabis perianth is only about 6 cells thick, so to distinguish calyx cells from corolla cells is best left to botanists. This book uses the botanically correct term bracts for the green or purple, resin-gland-studded specialized “leaves” encasing each female flower, and uses perianth or calyx for the translucent “veil” that clasps and covers about 60 to 90 percent of a mature seed.<\/p>\n\n\n\n

Hopefully, when growers use botanical terms such as calyx, bracts, stigma, pistil, anther and stamen, they will follow this book and use the terms correctly. Since readers will find seed catalogs and Internet sites calling bracts calyxes and stigmas pistils, it is important for readers to understand this confusion when reading other sources. Hopefully, this chapter will help get us all on the same page.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The stigmas on this ‘Haze’ \u00d7 ‘Northern Lights’ \u00d7 ‘Sensi Star’ flower top are just starting to turn a rusty color. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The perianth on this ‘Skunk #1’ can be seen near the top of the seed on the left. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The female perianth is clearly visible as a nearly transparent layer covering the seed. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Sexual Propagation<\/h2>\n\n\n\n

Sexual propagation is the process in which male and female sex cells (gametes) from separate parents unite in the female plant to form what will eventually mature into a new, genetically distinct individual. This process occurs when pollen from a male (staminate) parent unites with an ovule within the ovary of a female flower to create an embryo. This embryo, when mature and fully developed, will become a seed.<\/p>\n\n\n\n

In nature, cannabis is wind-pollinated. Male flowers shed pollen (dehiscence), dispersing millions of grains into the wind. Wind carries the pollen to a “chance” rendezvous and acceptance by a female stigma.<\/p>\n\n\n\n

Pollination occurs when male pollen grains land on a female stigma. The evolutionary attraction is both physical and chemical. The grain of pollen, with moisture found in the stigma, germinates. This is the best part: A grain of pollen germinates just like a seed, sending a taproot down, but instead of sending it into the ground, the grain of pollen sends the “root” down the stigma toward the ovary. Once united with the ovary, the pollen fertilizes the ovule. This union creates an embryo that grows within a seed coat. When mature in 4 to 6 weeks, the seed can be planted.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Fertilization occurs when the tiny grain of male pollen sticks to the stigma. Then it develops a tube through the style and releases 2 gametes, 1 to fertilize the ovule and 1 to fertilize the endosperm (double fertilization). Seeds are the result of this sexual propagation and contain genetic characteristics of both parents. Once fertilized with male pollen, female plants put the bulk of their energy into producing strong, viable seeds.<\/p>\n\n\n\n

Actual fertilization takes place when the minute grain of male pollen sticks to the stigma. The successful angiosperm pollen grain (gametophyte) containing the male gametes (sperm) gets transported to the stigma, where it germinates and its pollen tube grows down the style to the ovary. Its 2 gametes travel down the tube to where the gametophyte(s) containing the female gametes are held within the carpel. One nucleus fuses with the polar bodies to produce the endosperm tissues, and the other with the ovum to produce the embryo hence the term double fertilization.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Close-up photo of a female stigma shows that no resin glands are located on the entire length. The well-defined protruding growth appears as fuzz on the stalk of the stigma. The stigma is the plant version of a vagina. It is covered with stigmatic fluid that acts like glue when a piece of pollen lands on it. The fluid is packed with sugars that are food for the pollen. When in place, the grains of pollen start growing a new \u201cpollen tube,\u201d a long tunnel that pushes through the tissue of the style all the way to the ova, where it fuses with egg cells to create a little baby plant. The ovules go through a series of steps known as meiosis, a type of cell division by which a cell duplicates into 2 genetically identical daughter cells. The chromosomes in the cell nucleus separate into 2 matching sets of chromosomes, each with a nucleus.<\/p>\n\n\n\n

Deoxyribonucleic acid (DNA) or “genetic material”* is coiled into long strands or chromosomes. The DNA is located inside the nucleus of each cell. When cannabis is pollinated, each individual seed inherits 10 different chromosomes from the male, and 10 different chromosomes from the seed mother\u201420 chromosomes total. Each seed has 2 copies of each of the 10 chromosomes, or 1 full genome each. There are 2 copies of every gene in the plant, 1 from the mother and 1 from the father. Every cell in the plant has a copy of this unique DNA. The genetic code of this unique individual is embedded in a specific location along the length of the chromosome strands.<\/p>\n\n\n

\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Each seed<\/strong> contains genes from both parents. Progeny grown from seed usually have slightly different traits than other plants from the same seed lot. The same happens in humans; biological children are different from one another in many aspects and at the same time they resemble their parents. In cannabis, the variability is marked, like it is in apples.<\/p>\n\n\n\n

Sexual reproduction is used to cross different individuals with a population or family of plants. It can also be used to hybridize unrelated lines and inbreed their offspring. This phenomenon, \u201crecombination of traits,\u201d also gives breeders the opportunity to recover individuals with a combination of the positive traits of both parental lines.<\/p>\n\n\n\n

Genes are hereditary units that consist of a sequence of DNA that resides in a precise place on a chromosome and it determines a specific trait in cannabis. Little bits of DNA are codes or templates for proteins.<\/p>\n\n\n\n

Proteins are made in the sequence of DNA. Like instructions in a recipe for making brownies, the DNA and sequence of proteins is the recipe or instructions.<\/p>\n\n\n\n

Having 2<\/strong> versions of the same protein<\/strong>, from 2 different genes, is better than having just 1, particularly if the protein plays some vital part in cannabinoid production. This effect is called overdominance. For example, if there are 2 different proteins, and both work well but 1 works a little better under hot conditions and the other works well under cool conditions. Having 2 versions of the same protein gives the plant a wider range of climates where it produces effectively. See \u201cMultiline,\u201d on page 531.<\/p>\n\n\n\n

The majority of DNA is the same; it deals with basic cell processes, photosynthesis, chlorophyll production, etc. A few or a combination of genes control variables such as height, leaf shape, fragrance, and disease resistance. But we are not sure exactly which genes are responsible for specific traits, even though the cannabis genome has been mapped. These traits are influenced by multigene families (a group of genes that evolved to become a little bit different from each other, even though they started out as copies of the same gene). Knowing the named genes<\/strong> would make it easier to find individual plants with your desired traits. But single genes controlling specific traits of cannabis are neither isolated nor well studied.<\/p>\n\n\n\n

Multigene traits allow you to fine-tune to your favorite characteristics. For example, a single gene that controls leaf size would give only 2 leaf sizes, large and small. Many genes influencing the same trait provide many different leaf sizes.<\/p>\n\n\n\n

Naturally occurring mutated cannabis genes<\/strong> are uncommon. They are abnormal genes that are mutations of normal genes. When a mutated gene combines with a normal gene, there is no detrimental outcome. But when 2 mutated genes join, the result is much different.<\/p>\n\n\n\n

For example, in people and animals the number of albinos or dwarfs is minimal. The same is true in cannabis. Growing large populations of cannabis or treating cannabis with stress or chemicals will bring about mutation. Overall, most cannabis plants grow normally with no mutations whatsoever. Many different genes control the desirable traits we care about. Broken recessive genes do not play a role in most breeding programs.<\/p>\n\n\n\n

Deleterious recessive genes:<\/strong> The plants most likely to have the same dangerous mutation in the immediate family are inbred. Marry your sister and inbreeding genes take over to start all kinds of problems because recessive deleterious genes appear.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

‘Royal Moby’, predominantly sativa, is very THC-potent. This variety is quite similar to ‘Moby Dick’ from Dinafem Seeds in Spain. Once a good variety reaches the marketplace, many similar varieties appear within a year or two.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Classical Cannabis Breeding<\/h2>\n\n\n\n

Classical breeding is an ancient cyclical process that cannabis breeders still use today. Decisions are made based on observation of large numbers of plants; the breeder does not know exactly what genes have been introduced to the new cultivars. All the breeder can do is choose plants based upon visual inspection, smell, and gut-feeling.<\/p>\n\n\n\n

Classical cannabis breeding is simple: 2 varieties, a male and a female, are chosen. Each parent has desirable characteristics\u2014fragrance, potency, mold resistance, and so forth. The male pollen fertilizes the female flower and their genes combine into a new genetic mix contained in the seed.<\/p>\n\n\n\n

The next step is to choose individual plants with the desirable traits of both parents. A lot of the time you get lucky and the offspring carry the desirable genes and traits. Cannabis breeders often take clones of these desirable individual plants. Too often they do not retain the male plant and have a \u201cclone only\u201d variety.<\/p>\n\n\n\n

When a desirable trait has been bred into a plant, crossing other plants to this parent makes new plants similar to the favored parent. For example, to make the mildew-resistant progeny of the cross most like the high-yielding parent, the progeny will be backcrossed to that parent for several generations (see \u201cBackcrossing,\u201d page 519). This process removes most of the genetic contribution of the mildew-resistant parent.<\/p>\n\n\n\n

To breed for mold resistance, grow out plants in moldy conditions. Remove plants from garden that contract mold easily. Keep plants that do not get mold or are late to get mold. Breed plants that do not mold.<\/p>\n\n\n\n

It is very difficult to isolate specific genes to guarantee specific qualities such as extreme resistance to powdery mildew or insect and mite attacks. There are recessive genes and dominant genes that are controlled by alleles; this is where breeding becomes much more complex. See \u201cInfluence of Alleles.\u201d<\/p>\n\n\n\n

Other traits, such as acclimatization to a specific climate, are relatively easy because the plants that grow best in the environment are continually selected. Organic gardeners breed plants acclimated to their outdoor climate and achieve much higher yields. For this reason, organic medical gardeners in Northern California are able to grow 10-pound (4.5 kg) plants.<\/p>\n\n\n\n

To breed for cold tolerance, grow out plants in cold conditions. Remove plants that suffer cold damage easily. Breed plants that withstand cold temperatures.<\/p>\n\n\n\n


\n\n\n\n

Thin-Layer Chromatography<\/h3>\n\n\n\n

Thin-layer chromatography can be used to take tests and make selections based upon cannabinoid profiles. Cannabinoid profiles are similar throughout a plant\u2019s life stages, and breeding decisions can be based upon these profiles. For example, the cannabinoid profile can be tested on 2-month-old seedlings. Plants with desirable profiles are kept, and those with undesirable plants are culled.<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

See \u201cModern Cannabis Breeding\u201d at the end of this chapter for an overview of marker-assisted selection (MAS).<\/p>\n\n\n\n

<\/div>\n\n\n\n

Influence of Alleles<\/h2>\n\n\n\n

The phenotypes seen in a given individual are the result of an interaction between the plant\u2019s genotype and the environment. For example, here are 3 phenotypes: short, medium, and tall. Remember, the genotype describes the genetic condition responsible for the phenotype, and to represent it in discussion we assign it symbols.<\/p>\n\n\n\n

<\/div>\n\n\n\n
Phenotypes<\/strong><\/td>Genotypes<\/strong><\/td><\/tr>
short<\/td>ss<\/td><\/tr>
medium<\/td>ss<\/td><\/tr>
tall<\/td>SS<\/td><\/tr><\/tbody><\/table><\/figure>\n\n\n\n
<\/div>\n\n\n\n

There are always 2 versions of every gene (allele). For example, if there are 2 lower-case \u201cs\u201d plants with \u201csmall stature\u201d the plant will be shorter. But if the plant has capital \u201cS\u201d and has the \u201ctall\u201d gene, the phenotype is tall. If both genes are inherited, the plant is medium height.<\/p>\n\n\n\n

Homozygous \/ heterozygous: These are terms used in describing the genotypic condition of a plant, with regard to the similarity of the alleles for a given trait. If a plant is homozygous for a given trait, it has 2 copies of the same allele. If a plant is heterozygous, it has 2 different alleles for a given trait.<\/p>\n\n\n\n

The offspring inherits 1 set of alleles from each parent. This inheritance of alleles can be homozygous (both alleles are the same) or heterozygous (each allele is different). Furthermore, recessive alleles do not surface completely for several generations. The influence of alleles makes it impossible to use simple mathematical probability to predict the outcome of offspring.<\/p>\n\n\n\n

<\/div>\n\n\n\n
\n\n\n\n
<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

In August 2011, Dr. Kevin McKernan announced that his company had successfully mapped the genomes (shotgun sequence) for Cannabis sativa (‘Chemdawg’ variety) and later C. indica (‘LA Confidential’). Medicinal Genomics then publicized its work on C. sativa via Amazon\u2019s EC2, a cloud-computing service that gives free access to the scientific community. Search “Cannabis genome EC2 cloud” on www.google.com<\/a> for more information.<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This indica-dominant \u2018Peyote Purple\u2019 was developed by CannaBioGen.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Dominant and Recessive Traits<\/h2>\n\n\n\n

Dominant and recessive traits are dictated by alleles that are inherited from both parents. But, even though the cannabis genome has been decoded, specific gene function has not been deciphered. Consequently, the examples below must be used as guidelines because many traits are driven by a combination of genes.<\/p>\n\n\n\n

Dominant:<\/strong> An intra-allelic interaction such that the presence of an allele of 1 parent masks the presence of an allele from another parent plant, in the expression of a given trait in the progeny. Only the dominant trait is shown in the first generation of offspring. Of the F2 generation, 75 percent will also show the dominant condition.<\/p>\n\n\n\n

Recessive:<\/strong> An intra-allelic interaction such that an allele of 1 parent is masked by the presence of an allele from the other parent plant, in the expression of a given trait in the progeny. The recessive trait is not shown in the first generation of progeny (F1) but will reappear if siblings are mated, and the F2 progeny will result in 25 percent of plants showing the recessive condition.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

\u2018Afghani\u2019-dominant crosses are in the foreground and sativa-dominant crosses are in the background. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The \u2018Afghani\u2019 seedling on the left demonstrates the dominant traits of short, squat growth and broad leaves. The \u2018Kush\u2019 seedling on the right demonstrates the dominant traits of taller growth, and has much different leaf formation.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This genetically purple \u2018Mexican\u2019 sativa from 1980 is one of the varieties that added a purple hue to many current varieties. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n
\n\n\n\n

Numbers Can Be Misleading<\/strong><\/p>\n\n\n\n

Simplistic math models are often used to explain what happens when male and female cannabis plants are crossed. These models do not take into account dominant and recessive genes, they allow nothing for genotype and phenotype, and they do not consider that specific traits may be controlled by many different genes.<\/p>\n\n\n\n

Use the Punnett square to help predict the outcome.<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

We can demonstrate Mendel\u2019s Laws in a Punnett square to predict the possible outcome of 1 genetic trait (monohybrid). The example below uses Mendel\u2019s pea plant experiment. He crossed 2 yellow pea plants, and they produced 3 quarters yellow peas and 1 quarter green peas. The Punnett square accounts for all the possible combinations for 1 gene. Each of the 4 squares in the box represents 1 new offspring.<\/p>\n\n\n\n

A monohybrid is a genetic cross between 2 parents; 1 parent has 2 dominant alleles for a specific gene and the other 2 recessives for the same gene. The offspring, monohybrids, have 1 dominant and 1 recessive allele for that gene. A cross between the offspring produces a 3:1 ratio when grown out of the following (F2) generation of dominant:recessive phenotypes.*<\/p>\n\n\n\n

*This example is based on Mendel\u2019s monohybrid pea plant cross. Little information is available to the public about specific gene loci of cannabis. An excellent worksheet is available at BiologyCorner.com (www.biologycorner.com\/worksheets\/pennygene_key.html<\/a>).<\/p>\n\n\n\n

A dihybrid is a genetic cross between parents that have 2 characteristics controlled by genes at different loci.* The example below shows both parents AaBb and AaBb are heterozygous for both traits (color and height). The possible outcome can be predicted with the help of a Punnett square. The example below is based on Mendel\u2019s dihybrid cross with pea plants. Note the dihybrid Punnett square has 16 different possibilities. This will yield a genotype ratio of 1:2:1:2:4:2:1:2:1 and a phenotype ratio of 9:3:3:1.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

The monohybrid cross ratio is demonstrated in the Punnett square.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

This example is based on Mendel\u2019s pea plant cross, and is hypothetical because little information is available to the public about specific gene loci of cannabis.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

At 4 weeks of flowering, \u2018Big Bud\u2019 is just starting to put on weight. The genetics of \u2018Big Bud\u2019, from northwestern USA, are found in many of today\u2019s varieties.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

True Breed<\/h2>\n\n\n\n

True-breeding (or inbred) plants make the best breeding stock. A true-breeding variety (aka inbred line [IBL] and true line) of cannabis is the result of a seed lot that has been bred for generations while selecting repeatedly for specific traits. These plants breed true for the specific traits\u2014cannabinoid content, strong growth, desirable aroma and flavor, and so on. There is little or no variation of traits, growth is uniform, and the outcome of future generations is easier to predict. These plants are said to be stable.<\/p>\n\n\n\n

If seed is true breeding (IBL) it should be easy to reproduce by open pollination. See \u201cMaking Seeds at Home,\u201d page 523, for a list of true-breeding seed varieties. After many generations of repetitive breeding and selecting for the same traits, other genetic traits may deteriorate. See \u201cInbreeding,\u201d page 518, for more information.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Hybrids<\/h2>\n\n\n\n

Hybrids are a product of a cross between genetically distinct parents. Hybrids do not reproduce their parent characteristics completely or reliably when reproduced sexually. Develop hybrid cultivars by using inbred (true-breeding) genetics or by segregating populations, inbreeding them, and selecting for hybrid line production. Various hybrid varieties are described below.<\/p>\n\n\n\n

<\/div>\n\n\n\n
\n\n\n\n

Hybrid Crosses<\/h3>\n\n\n\n

F1 hybrids (\u2018Northern Lights\u2019 \u00d7 \u2018Blueberry\u2019, \u2018Northern Lights\u2019 \u00d7 \u2018Haze\u2019)<\/p>\n\n\n\n

3-way crosses (\u2018Skunk #1\u2019\u2014a cross of (\u2018Mexican\u2019 \u00d7 \u2018Colombian\u2019) \u00d7 \u2018Afghani\u2019) 4-way or double cross hybrids (cross of 2 unrelated F1 hybrids \u2018Haze\u2019 (\u2018Afghani\u2019 \u00d7 \u2018Thai\u2019) \u00d7 (\u2018Mexican\u2019 \u00d7 \u2018Colombian\u2019).<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

F1 Hybrid Varieties<\/h3>\n\n\n\n

Achieve an F1 hybrid population by crossing 2 unrelated, true-breeding varieties. F1 hybrids are uniform when grown from seed, but, like all hybrids, they are genetically unstable. If reproduced sexually by inbreeding within the F1 population, the subsequent generation will be neither uniform nor similar to the F1 generation. Seeds from F1 hybrids, F2, will all be different; parents are heterozygous and diverse. They average twice as many monozygous genes and will be less vigorous. Seeds lose uniformity and hybrid vigor.<\/p>\n\n\n\n

F1 hybrid vigor<\/strong> (heterosis) occurs when 2 true-breeding varieties are crossed. The resulting seed and plant has \u201chybrid vigor,\u201d\u2014growing more robustly, stronger, faster, and with a heavier harvest than both parents. For example, a (\u2018Skunk#1\u2019 \u00d7 \u2018Blueberry\u2019) F1 hybrid grows faster and yields more than either the pure \u2018Skunk #1\u2019 or \u2018Blueberry\u2019 parent populations.<\/p>\n\n\n\n

F1 hybrid seeds grow into strong plants that produce well. Seed companies sell them because they keep clients coming back for more seeds every year. Few seed companies are interested in selling seed that is easy to open pollinate or easy to reproduce. Most seed companies make and release hybrids. Often \u201ctrue-breeding\u201d (IBL) seeds are not stable. Competitors in the unregulated market continually pirate seed varieties from companies that have developed the seeds.<\/p>\n\n\n\n

F1 hybrid is a term used in genetics and selective breeding. F1 stands for Filial 1, the first filial generation of seeds\/plants or animals.<\/p>\n\n\n\n

Maintaining strong, reliable F1 hybrid stock requires growing 2 different lines of the same parents. The 2 lines are crossed to make seed. Deviation is minimized. This is necessary to maintain genetic diversity and avoid recessive trait domination.<\/p>\n\n\n\n

For information on F2, F3, etc., generations, see “Filial Breeding,” on page 520.<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

Cross-pollination<\/strong> is the transfer of pollen from an individual male cannabis flower to a stigma of the flower of a female plant. Strive to combine the best qualities of a true-breeding line with other lines to increase desired traits when using cross-pollination in cannabis breeding.<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

Inbreeding<\/h2>\n\n\n\n

Inbreeding is crossing a group, family, or variety of plants within themselves with no addition of genetic material from an outside or unrelated population. \u201cTrue Breed,\u201d above, is an example of inbreeding. The most severe form of inbreeding is the self-cross (see \u201cSelf-Pollination (Selfing)\u201d on page 520), in which only one individual\u2019s genetic material forms the basis of subsequent generations. And 1:1 hybrid populations are only slightly less narrow, derived from the genetic material of 2 individuals. Such tight or narrow breeding populations lead to a condition called \u201cinbreeding depression\u201d upon repeated self-breeding or inbreeding.<\/p>\n\n\n\n

Inbreeding depression is a reduction in vigor (or any other character) due to prolonged inbreeding. This can manifest as a reduction in potency or a decrease in yield or rate of growth. Progress of depression is dependent, in part, on the breeding system of the crop.<\/p>\n\n\n\n

Cannabis is an outcrossing or cross-pollinating species. Cross-pollinated cannabis exhibits a higher degree of inbreeding depression when “selfed,” or inbred, than do selfing crops such as tomatoes, which can be selfed for 20 generations with no apparent loss in vigor or yield.<\/p>\n\n\n\n

In cross-pollinated crops, deleterious genes remain hidden within populations, and the negative attributes of these recessive traits can be revealed when inbred several generations.<\/p>\n\n\n\n

Inbreeding depression can be noticeable in S1 (see \u201cSelf-Pollinating (Selfing),\u201d on page 520) populations after a single generation of self-fertilization. When breeding cannabis using small populations, as is often the case with continual 1:1 mating schemes, inbreeding depression typically becomes apparent within three to four generations. This 1:1 breeding model is being used today by most commercial seed banks.<\/p>\n\n\n\n

Cannabis is naturally an outcrossing or cross-pollinating species and existed in wild breeding populations of hundreds if not thousands of individuals. Within these many individuals lies a wide range of versions of different genes. When only 1 or 2 plants are selected from this vast array as the breeding population, there is a drastic reduction in the genetic variability found in the original population, resulting in a genetic bottleneck. Once this variability is lost from a population, it is gone.<\/p>\n\n\n\n

If space is available, one way to overcome this problem is by maintaining 2 separate parallel breeding lines. After generations of inbreeding, when each of the inbred lines or selfed populations starts showing inbreeding depression, they are hybridized or outcrossed to each other to restore vigor and eliminate inbreeding depression while preserving the genetic stability of the traits under selection.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

‘Afghani #1’ is an inbred variety that was developed over years of inbreeding among large populations of cannabis plants. This landrace variety was collected in the mountains of Afghanistan by Sacred Seeds in the 1970s. It was one of the first pure indica true-breeding strains used in worldwide cannabis breeding programs. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Outbreeding<\/h2>\n\n\n\n

Outbreeding is the process of crossing or hybridizing plants or groups of plants with others to which there is no relation or the relationship is very distant. Any time a breeder is hybridizing using plants that reside outside of the family, group, or variety, hybrid seed is produced.<\/p>\n\n\n\n

An F1 hybrid seed is the first generation offspring that results from crossing 2 distinct true-breeding plants or populations. Each of the (true-breeding) parents was hybridized (outcrossed to each other) to produce the new generation, which is now comprised of genetics from both parental populations. Outcrossing results in the introduction of new and different genetic material to each of the pools.<\/p>\n\n\n\n

Often there are recessive genes or dominant genes that have not had a chance to fully express themselves until being inbred 3 to 4 times or being outcrossed several times. It is important to remember that cannabis is naturally outcrossing. Inbreeding plants that are naturally outcrossing too many generations sacrifice the health of offspring. Keep outcrossing plants healthy, and maintain diversity by growing large numbers of plants.<\/p>\n\n\n\n

Filial Breeding:<\/strong> Siblings of the same progeny lot and generation are interbred to produce new generations. The first hybrid generation of 2 distinct true-breeding lines is the F1 generation (F, filial). If 2 F1 siblings are bred, or the F1 population is allowed to be open-pollinated, the resulting generation is labeled F2, F3, F4, etc. Mating siblings chosen from the F2 results in the F3 population. F4, F5, F6 generations, etc., are produced the same way, by crossing plants of the same generation and progeny lot.<\/p>\n\n\n\n

Note:<\/strong> As long as any number of siblings of a generation (F[n]) are mated, the resulting generation is denoted (F[n+1]).<\/p>\n\n\n\n

Filial inbreeding with selection for specific traits is the most common method for establishing a pure or a true-breeding population in cannabis. After a few generations, genetic problems will surface.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

‘Super Skunk’ from Sensi Seeds is a good example of an outbred variety. It takes the stable ‘Skunk #1’ and adds a dominant amount of ‘Afghani’ genetics.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Backcrossing<\/h2>\n\n\n\n

Backcrossing is the pollination of a female flower from a male flower on the same plant. Backcross breeding is repeated crossing of progeny with one of the original parental genotypes. It is cross-pollinating 1 generation back to a previous generation; most often the progeny is crossed with the mother plant. The parent is called the \u201crecurrent parent.\u201d The nonrecurrent parent is the \u201cdonor parent.\u201d<\/p>\n\n\n\n

Backcross breeding is the most widely used form of breeding cannabis to date.<\/strong> Backcross breeding is simple and can be done with small populations of plants. Most often, the goal of backcrossing is to create a population from the genetics of a single parent (the recurrent parent).<\/p>\n\n\n\n

Donor parents are chosen based on desirable traits. One parent is backcrossed to another to introgress genetics, to move genes from one variety to another. Repeated backcrossing is the best way to achieve this goal.<\/p>\n\n\n\n

Use backcrossing to add desirable traits to a mostly ideal, relatively true-breeding genotype. The recurrent parent should be an ideal genotype such as an existing inbred line. Look for traits that are easily identified in the progeny of each generation. The best donor parent must possess the desired trait but should not be seriously deficient in other traits. Backcross line production is repeatable if the same parents are used. <\/p>\n\n\n\n

Simple backcrossing to incorporate a dominant trait is easy and very common among the vast majority of breeders. However, plants come with both dominant and recessive genes. When making selections, it is best to select for one single quality at a time and select the absolute best males possible from each population.<\/p>\n\n\n\n

Backcrossing also uses the terms “squaring” (to denote the second backcross to the same parent) and “cubing” (to designate the third backcross).<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

‘Apollo 13 BX’ (Back Cross) from TGA Genetics is a good example of a backcrossed variety.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Backcrossing: Incorporating a Dominant Trait<\/h3>\n\n\n\n

Step One:<\/strong> Cross recurrent parent \u00d7 donor parent<\/p>\n\n\n\n

Step Two:<\/strong> Select desirable plants showing the dominant trait, and hybridize selected plants to the recurrent parent. The generation produced is denoted BC1 (some cannabis breeders call this generation B\u00d71 [BC1= B\u00d71]).<\/p>\n\n\n\n

Step Three:<\/strong> Select plants from BC1 and hybridize with the recurrent parent; the resulting generation is denoted BC2.<\/p>\n\n\n\n

Step Four:<\/strong> Select plants from BC2 and hybridize with the recurrent parent; the resulting generation is denoted BC3.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Backcrossing: Incorporating a Recessive Trait<\/h3>\n\n\n\n

Recessive traits are more difficult to select for in backcross breeding since their expression is masked by dominance in each backcross to the recurrent parent. An additional round of open pollination or sib-mating is needed after each backcross generation to expose homozygous-recessive plants. Individuals showing the recessive condition are selected from F2 segregating generations and backcrossed to the recurrent parent as in the \u201cBackcrossing: Incorporating a Dominant Trait\u201d above.<\/p>\n\n\n\n

Step One:<\/strong> Cross recurrent parent \u00d7 donor F1 hybrid generation.<\/p>\n\n\n\n

Step Two:<\/strong> Select desirable plants and create an F2 population via full cross-pollination.<\/p>\n\n\n\n

Step Three:<\/strong> Select plants showing the desired recessive trait in the F2 generation. Hybridize selected F2-recessive plants to the recurrent parent. The generation produced is denoted BC1.<\/p>\n\n\n\n

Step Four:<\/strong> Select plants from BC1 and create a generation of F2 plants via sib-mating; the resulting generation can be denoted BC1F2.<\/p>\n\n\n\n

Step Five:<\/strong> Select desirable BC1F2 plants showing the recessive condition, and hybridize with the recurrent parent; the resulting generation is denoted BC2.<\/p>\n\n\n\n

Step Six:<\/strong> Select plants from BC2, and create an F2 population via sib-mating; denote the resulting generation BC2F2.<\/p>\n\n\n\n

Step Seven:<\/strong> Select plants showing the recessive condition from the BC2F2 generation, and hybridize to the recurrent parent; the resulting generation is denoted BC3.<\/p>\n\n\n\n

Step Eight:<\/strong> Grow out BC3, select and sib-mate the most ideal candidates to create an F2 population, where plants showing the recessive condition are then selected and used as a basis for a new inbred, or open-pollinated, seed line.<\/p>\n\n\n\n

This new generation created from the F2 is a population that consists of, on average, about 93.7 percent of genes from the recurrent parent, and only about 6.3 percent of genes leftover from the donor parent.<\/p>\n\n\n\n

Note:<\/strong> Only homozygous-recessives were chosen for mating in the BC3F2 generation; the entire resulting BC3F3 generation is homozygous for the recessive trait and breeds true for this recessive trait. This new population meets the breeding objective, consisting mainly of genetics from the recurrent parent while breeding true for the introgressed recessive trait.<\/p>\n\n\n\n

Backcross-derived lines are most often well-adapted to the environment in which they will be grown. Indoor garden rooms are easily replicated, and plants grow in a similar environment to that in which they were bred. Progeny therefore need less extensive environmental field-testing.<\/p>\n\n\n\n

If 2 or more characters are to be introgressed into a new seed line, these would usually be tracked in separate backcross programs, and the individual products would be combined in a final set of crosses after the new populations have been created by backcrossing.<\/p>\n\n\n\n

Backcrossing has some drawbacks to consider. If the recurrent parent is not a very true-breeding variety, the resulting backcross generations may segregate, and many of the desirable traits intended to be reproduced in the line may fail to be replicated reliably. Additionally, the limitation of backcrossing is that the “improved” variety differs only slightly from the recurrent parent, typically focusing on one specific trait. If the goal is to introgress multiple traits into a new population, other breeding techniques like inbreeding or recurrent selection might be more effective and rewarding.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Small male flower appears on this Chemdog inbred female flower. Inbreeding often causes male intersex flowers to form on female plants.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Self-Pollinating (Selfing)<\/h2>\n\n\n\n
\n\n\n\n
<\/div>\n\n\n\n

“Clone Only\u201d Varieties Often, 2 hybrid plants are crossed and the \u201cvariety\u201d is given a name, but soon the male plant is lost, and the plant is available only as a clone. In this case the plant must be \u201cselfed\u201d to produce male flowers on a female plant.”<\/p>\n\n\n\n


\n\n\n\n
<\/div>\n\n\n\n

Self-pollinating (aka selfing) is the process of creating seed by pollinating a plant with its own pollen. Self-pollinating is a plant having sex with itself. Self-crossing can derive populations of plants from a single individual. The first generation population derived from selfing an individual is called the S1 population. An individual from S1 that is selfed again is called S2. Subsequent generations derived in the same manner are denoted S3, S4, etc.<\/p>\n\n\n\n

Traits for which the plant is homozygous remain homozygous upon selfing, whereas heterozygous loci segregate and may demonstrate novel expressions of these characters.<\/p>\n\n\n\n

We know homozygous loci remain homozygous in future generations upon selfing. Heterozygous loci are increased by 50 percent. Every subsequent generation will be 50 percent more homozygous than the parent from which it was derived.<\/p>\n\n\n\n

Repeated selfing, or single-seed descent, is the fastest way to achieve homozygosity within a group or family. The more plants grown from a selfed population, the better probability a breeder has of finding selfed progeny that show all the desired traits.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Self-Pollinating Breeding<\/strong><\/h3>\n\n\n\n

Step One:<\/strong> Identify superior genotypes for the trait under selection.<\/p>\n\n\n\n

Step Two:<\/strong> Cross superior genotypes and select improved progeny.<\/p>\n\n\n\n

Step Three:<\/strong> Repeat steps One and Two over a series of generations.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Feminized Seeds<\/h2>\n\n\n\n

Breeders produce all-female or feminized seeds by obtaining pollen from a male flower on an individual female and using this pollen to fertilize another female plant. In order to produce feminized seed, pollen must be collected from male flowers that occur on an otherwise female plant.<\/p>\n\n\n\n

As stated in “Sexual Propagation,” there are 20 chromosomes in each plant cell. Female cannabis plants have 2 copies of the X chromosome, and their genotype is expressed as XX. Male plants have 1 copy of the X chromosome and 1 copy of the Y chromosome. The genotype expression of male plants’ sex chromosomes is XY. However, the ability to change sex based on external factors is believed to be controlled by X-autosome or X-A.<\/p>\n\n\n\n

When pollen is created within the plant, 1 of each of the sets of chromosomes from each parent is packaged into the cells that develop into pollen. This process, mitosis, is the process of a single cell dividing into 2 identical cells. Each cell contains the same chromosomes and genetic content as the parent. Each pollen grain or ovule contains 10 chromosomes, 1 copy of each pair. When the pollen deposits the genetic material into the ovule, the 10 chromosomes from the pollen and the ovule unite to make a total of 20 chromosomes, a full genetic complement. There are 2 gametes: 1 fertilizes the ovule of the female, the other fertilizes the endosperm of the seed that gives the new plant its initial food source and chemical profile.<\/p>\n\n\n\n

Use a Punnett Square to predict the outcome of a breeding cross. It is used to determine the probability of an off- spring having a particular genotype. The Punnett square summarizes all possible combinations of 1 maternal allele with 1 paternal allele.<\/p>\n\n\n\n

The next step is to produce male flowers on a female plant. Pollen from the male flowers is used to fertilize the same female plant (selfing) or other female plants (outbreeding). If this female plant is the same self-pollinated plant, it will be a mix of its own genes. If the plant is different and from outside the genetic pool of the first plant, new genes will be introduced to the outbred cross.<\/p>\n\n\n\n

There are 2 common ways to induce male flowers on a female plant\u2014 inducing environmental stress and altering hormone levels.<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Seeds must develop completely before they are ready to harvest. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

\u2018Sour Bubbly\u2019 is one of the many new femi- nized autoflowering varieties.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Punnett square view of a simple male:female cross<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This Punnett square of a female cross demonstrates that female:female breeding crosses produce only female (XX) offspring. Pollen from a few male flowers that appear on a female plant is used to produce female only (XX) offspring.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Environmental Stress<\/h3>\n\n\n\n

Pollen from environmentally stressed intersex plants is used to fertilize females that will produce mostly (more than 99%) female seeds. But, the offspring will have intersex tendencies. More pronounced intersex tendencies in parent plants bring about the same traits in offspring. There is no way around this simple genetic fact.<\/p>\n\n\n\n

Produce feminized seed by collecting pollen from carefully selected, latent, stress-induced male flowers on a female plant, and use it to pollinate female plants.<\/p>\n\n\n\n

One of the easiest stress techniques is to let the plant flower and continue growing. Sooner or later most females will produce a few male flowers. Other stress methods include irregular light cycle or low temperature. The process is time-consuming but yields mostly female plants when grown without stress. As soon as stress is introduced, intersex tendencies can quickly surface.<\/p>\n\n\n\n

Remember, genetic weaknesses\u2014 including a predisposition for male flowers to appear on female plants\u2014will be passed to offspring. To select against the intersex condition, take female breeding candidates and grow them under stressful conditions such as irregular light cycle or high heat. Only plants that resist intersexuality under stressful conditions should be tested to produce an all-female seed lines. Intersex-resistant plants are called \u201ctrue females.\u201d Always select females that resist changing sex to pass the trait to offspring.<\/p>\n\n\n\n

Environmental stress is not the only way to produce male flowers on a female plant.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Alter Hormone Levels<\/h3>\n\n\n\n

Feminized seeds can also be produced with applications of sprays that alter the hormonal concentration of the plant. There are 2 popular methods of inducing male flowers on a female plant. Both methods require a spray application to the plant.<\/p>\n\n\n\n

Colloidal silver (CS) consists of very small silver particles suspended in water. There is quite a bit of disagreement as to the overall safety and efficacy of colloidal silver. Please research safety aspects before using it. In cannabis, CS inhibits female flowering hormones, so male hormones dominate. Male flowers appear on an otherwise female plant. Purchase high-quality CS at health food stores, natural pharmacies, or via Internet stores. Look for CS that is 15 to 30 ppm concentration. Or you can make it yourself at home, which is a bit more complicated. There are many YouTube videos and Internet methods to follow. They involve using a 9- to 12-volt battery or generator that can produce at least 250 milliamperes, a silver rod or coin, and distilled water. For instructions on colloidal silver, see chapter 22, Additives.<\/p>\n\n\n\n

Silver thiosulfate<\/strong> is more difficult to find than CS, but it is available on the Internet. It works on the same principles as CS. Check package for spray concentration and application.<\/p>\n\n\n\n

Gibberellic acid (GA3)<\/strong> is a very popular hormone used to produce feminized seeds. See chapter 22, Additives, for complete information.<\/p>\n\n\n\n

Ethylene<\/strong>, a plant hormone, plays a major role in the determination of sex by regulating which flowers should be produced\u2013male or female. High concentrations of ethylene sprayed on male flowering plants causes female flowers to form. See chapter 22, Additives, for more information on ethylene.<\/p>\n\n\n\n

Applying ethylene-inhibiting agents to pistillate individuals as they enter flowering results in the formation of stamens in place of pistils. Breeders use this technique to create \u201cfeminized\u201d seeds (all-female [gynoecious] seed lots).<\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

Large, round, capitate-stalked resin gland heads add perspective to minute grains of male pollen. (MF)<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

This autoflowering feminized \u2018Big Low\u2019 from Seeds of Life is one example of the many new \u201cauto-fems\u201d available today.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n
\n
\"\"<\/figure><\/div>\n\n\n
<\/div>\n\n\n\n

\u2018Jack 47\u2019 from Sweet Seeds proves that autoflowering feminized seeds produce plenty of resin.<\/em><\/p>\n\n\n\n

<\/div>\n\n\n\n

Day-Neutral (aka Autoflowering) Feminized Varieties<\/h2>\n\n\n\n

Day-neutral cannabis plants are called autoflowering plants in the popular cannabis industry today. Autoflowering feminized seeds are produced the same way as other seeds, except 1 or both of the parents are autoflowering. These varieties were originally created by cross-pollinating Cannabis ruderalis that flowers after 3 to 4 weeks of growth with regular cannabis plants that flower under 11 to 12 hours of darkness after the vegetative growth stage is complete. Autoflowering varieties are relatively easy to create, but they take time to stabilize. Seed companies sell them to you every year so that you do not have to reproduce them.<\/p>\n\n\n\n

Breeding new autoflowering varieties is more difficult when making hybrids and incorporating a non-autoflowering variety. A few short-day cannabis varieties are heterogeneous, contain the recessive day-neutral (autoflowering) genes and dominant short-day genes.<\/p>\n\n\n\n

Autoflowering plants have homozygous recessive traits for the day-neutral genes. Most short-day varieties and autoflowering crosses produce few autoflowering progeny in the F1 generation. When recrossed, the F2 generation contains approximately 25 percent of homozygous recessive plants that are autoflowering. Once this point is reached, the population requires further stabilization.<\/p>\n\n\n\n

Learn to breed with regular seeds before starting to breed autoflowering feminized plants. Once you have a good background with the variety you are breeding, you can start autoflowering breeding. Remember, half the equation is finding a male that fits your needs and is super strong-smelling, resinous from an early age, disease- and pest-resistant, and so on. Such males are often short and squat, with female-like growth and stature characteristics.<\/p>\n\n\n\n

Once autoflowering plants are genetically stable, they are feminized via inbreeding pollen from a male flower that occurs on a predominantly female plant. The process is fairly simple and is outlined below.<\/p>\n\n\n\n

<\/div>\n\n\n\n

Breeding Basics for Autoflowering Feminized Plants<\/strong><\/p>\n\n\n\n

    \n
  1. Cross regular seed with an autoflowering variety.<\/li>\n\n\n\n
  2. Cross F1 seeds together again under a 12\/12 photoperiod.<\/li>\n\n\n\n
  3. The first generation contains recessive autoflowering genes that manifest in the next generation.<\/li>\n\n\n\n
  4. The next generation has some autoflowering genes.<\/li>\n\n\n\n
  5. Plant and stabilize over time to get 100 percent autoflowering plants.<\/li>\n\n\n\n
  6. Feminize 100 percent autoflowering female and produce seeds.<\/li>\n<\/ol>\n\n\n\n

    Backcross regular seeds to an autoflowering male to achieve an autoflowering plant too.<\/p>\n\n\n\n

    Breeding new autoflowering varieties is more difficult when making hybrids and incorporating a non-autoflowering variety. A few short-day cannabis varieties are heterogeneous, contain the recessive day-neutral (autoflowering) genes and dominant short-day genes.<\/p>\n\n\n\n

    Autoflowering plants have homozygous recessive traits for the day-neutral genes. Most short-day varieties and autoflowering crosses produce few autoflowering progeny in the F1 generation. When recrossed, the F2 generation contains approximately 25 percent of homozygous recessive plants that are autoflowering. Once this point is reached, the population requires further stabilization.<\/p>\n\n\n\n

    Learn to breed with regular seeds before starting to breed autoflowering feminized plants. Once you have a good background with the variety you are breeding, you can start autoflowering breeding. Remember, half the equation is finding a male that fits your needs and is super strong-smelling, resinous from an early age, disease- and pest-resistant, and so on. Such males are often short and squat, with female-like growth and stature characteristics.<\/p>\n\n\n\n

    Once autoflowering plants are genetically stable, they are feminized via inbreeding pollen from a male flower that occurs on a predominantly female plant. The process is fairly simple and is outlined below.<\/p>\n\n\n\n

    Breeding Basics for Autoflowering Feminized Plants<\/p>\n\n\n\n

      \n
    1. Cross regular seed with an autoflowering variety.<\/li>\n\n\n\n
    2. Cross F1 seeds together again under a 12\/12 photoperiod.<\/li>\n\n\n\n
    3. The first generation contains recessive autoflowering genes that manifest in the next generation.<\/li>\n\n\n\n
    4. The next generation has some autoflowering genes.<\/li>\n\n\n\n
    5. Plant and stabilize over time to get 100 percent autoflowering plants.<\/li>\n\n\n\n
    6. Feminize 100 percent autoflowering female and produce seeds.<\/li>\n<\/ol>\n\n\n\n
      <\/div>\n\n\n\n

      Backcross regular seeds to an autoflowering male to achieve an autoflowering plant too.<\/h2>\n\n\n\n

      Making seeds at home requires a secure garden area that can be dedicated to breeding. It must be clean and free of male pollen. Indoor garden rooms and greenhouses are best suited to breeding cannabis if there is danger of rogue pollen from neighboring male cannabis plants contaminating faraway seed crops. If growing more than 1 male to produce pollen, measures must be taken to isolate each male once they start producing pollen. Segregate pollen-producing male plants by keeping them in an enclosed area as far away from flowering females as possible.<\/p>\n\n\n\n

      The basics of breeding cannabis at home are simple:<\/strong><\/p>\n\n\n\n

        \n
      1. Start with a diverse group of plants and know their traits.<\/li>\n\n\n\n
      2. Cross desirable plants.<\/li>\n\n\n\n
      3. Select desirable individuals and maintain diversity.<\/li>\n<\/ol>\n\n\n\n

        Slow and steady breeding is not dynamic, but it is very effective. Pollinate a few branches on several plants and have more seeds than you can grow. The breeding project is never finished!<\/p>\n\n\n\n

        Grow small seed crops inside a garden closet or inside a portable wardrobe made from fine cloth or something similar that will block pollen from entering or escaping.<\/p>\n\n\n\n

        Recordkeeping is the most important aspect of cannabis plant breeding. Accurate written and photographic records with dates help you to make informed decisions.<\/p>\n\n\n\n

        <\/div>\n\n\n\n
        \n\n\n\n

        Naming Varieties<\/h3>\n\n\n\n

        Many varieties have names known only in the local area where they are cultivated. These varieties often bear the name of the region from which they originated or a person with whom they are associated. Other varieties are named for fragrance, bud size, effect, or appearance. The names are often very descriptive and creative!<\/p>\n\n\n\n


        \n\n\n\n
        <\/div>\n\n\n\n

        Making Seeds: Step-by-Step<\/h2>\n\n\n\n

        Step One: What is your goal?<\/strong> Common goals include making seeds for next year\u2019s crop, reproducing new parents just like the old ones, and adding new traits to existing plants. When setting a goal, work with a single trait at a time so that it is easier to control the selection and outcome. Breeding will bring about dominant and recessive genes. Each has its own qualities. Impeccable recordkeeping and a consistent, stable climate are essential. Write out a breeding plan and plot your findings.<\/p>\n\n\n\n

        <\/div>\n\n\n
        \n
        \"\"<\/figure><\/div>\n\n\n
        <\/div>\n\n\n\n

        \u2018Skunk #1\u2019 is a stable breeding plant and is at the base of many varieties and breeding programs. Choose stable breeding stock whenever possible. (MF)<\/em><\/p>\n\n\n\n

        <\/div>\n\n\n\n

        Step Two: Select parents\u2014(A) female, (B) male\u2014to breed.<\/strong> Few varieties commercially available to medical cannabis gardeners are true-breeding and stable. Most often seeds have many different genes and are not stable or uniform. True-breeding seed stock ensures F1 hybrid seeds that have hybrid vigor. Stable true-breeding stock is sometimes difficult to find from seed companies, and gardeners often elect to stabilize their own.<\/p>\n\n\n\n

        Simply crossing 1 of the stable true-breeding varieties (see list below) with another stable variety will produce an F1 hybrid.<\/p>\n\n\n\n

        List of relatively stable seeds:<\/p>\n\n\n\n